Está en la página 1de 71

Fuel Processing Technology, 34 (1993) 1 71

E l s e v i e r S c i e n c e P u b l i s h e r s B.V., A m s t e r d a m

Review

C o m b u s t i o n o f c o a l as a s o u r c e o f N 2 0 e m i s s i o n

M.A.

Wbjtowicz,

J.R. Pels

and

J.A.

Moulijn*

Department of Chemical Engineering, Delft University of Technology, Julianalaan 136,


2628 BL Delft (The Netherlands)
( R e c e i v e d O c t o b e r 8th, 1992; a c c e p t e d J a n u a r y 25th, 1993)

TABLE

OF CONTENTS

Abstract

..............................................

1. I n t r o d u c t i o n

..........................................

2. N 2 0 a n d t h e e n v i r o n m e n t
..................................
2.1 G r e e n h o u s e effect a n d o z o n e l a y e r d e p l e t i o n . . . . . . . . . . . . . . . . . . . . .
2.2 C o n t r i b u t i o n of N 2 0 to t h e g r e e n h o u s e effect . . . . . . . . . . . . . . . . . . . .
2.3 I n v e n t o r y o f N 2 0 e m i s s i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4
4
7
9

3. N 2 0 C h e m i s t r y in c o m b u s t i o n of c o a l . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1 D e v o l a t i l i s a t i o n a n d p y r o l y s i s
.............................
3.2 G a s - p h a s e r e a c t i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 E q u i l i b r i u m c o n s i d e r a t i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.2 G a s - p h a s e N 2 0 / N O f o r m a t i o n a n d d e s t r u c t i o n . . . . . . . . . . . . . . . . .
3.3 H e t e r o g e n e o u s r e a c t i o n s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1 H e t e r o g e n e o u s N 2 0 / N O f o r m a t i o n a n d d e s t r u c t i o n o n c h a r s u r f a c e . . . .
3.3.2 H e t e r o g e n e o u s N 2 0 / N O f o r m a t i o n a n d d e s t r u c t i o n o n c a t a l y t i c s u r f a c e s .

13
19
23
23
26
34
34
36

4. N 2 0 / N O F o r m a t i o n a n d d e s t r u c t i o n in f l u i d i s e d - b e d c o m b u s t o r s
4.1 E f f e c t o f f u e l t y p e a n d o p e r a t i n g c o n d i t i o n s
.....................
4.1.1 T e m p e r a t u r e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.2 A i r s t a g i n g a n d e x c e s s a i r . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.3 F u e l t y p e . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.4 L i m e s t o n e a d d i t i o n . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.5 I n j e c t i o n o f a n a d d i t i v e (NH3, u r e a a n d c y a n u r i c acid)
4.2 F u e l - N c o n v e r s i o n to N 2 0 a n d N O . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3 N 2 0 A b a t e m e n t s t r a g e t i e s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

43
43
43
43
44
48
49
51
57

5. C o n c l u d i n g r e m a r k s
References

......................................

.............................................

* T o w h o m c o r r e s p o n d e n c e s h o u l d be a d d r e s s e d .

0378-3820/93/$06.00

,~ 1993 E l s e v i e r S c i e n c e P u b l i s h e r s B.V.

...........

...........

59
62

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

Abstract

Coal combustion is examined as a source of nitrous oxide pollution and a review of relevant
research is presented. The role N20 plays in global warming and in stratospheric ozone depletion
is explained and comparisons are made between N20 and other greenhouse gases. A balance of
known N20 sources and sinks is also reported. Nitrous oxide emerges as a powerful and long-lived
greenhouse gas with atmospheric concentration increasing at an annual rate of 0.3%. Most
anthropogenic sources of N20 are of comparable strength and all should be subject to control. The
fate of fuel-bound nitrogen is discussed in detail starting with coal devolatilisation and pyrolysis,
through gas-phase and heterogeneous N20/NO~ formation and destruction mechanisms, to conclude with the relevant side reactions. Kinetic data are provided where available and compiled in
tables. The main nitrogenous products of coal pyrolysis are HCN and NH3, and both can act as
gas-phase N20/NO precursors. Nitrous oxide is preferentially formed from cyano species, whereas
NH3-based compounds tend to react mainly to NO. Gas equilibrium calculations show that N20
concentrations in flue gas are several orders of magnitude above their equilibrium values.
Gas-phase formation of N20 is competitive with respect to NO formation. As temperature decreases, more N20 is formed at the expense of NO. Heterogeneous N20/NO formation probably
involves a similar trade-off and underlying mechanisms are discussed. Only up to 10% of charbound nitrogen has been found to form N20. Destruction mechanisms of N20/NO on char surface
are important under fluidised-bed combustion (FBC) conditions, especially in the presence of CO.
NO reduction on char is believed to be a negligible source of N20. Temperature has been identified
as the most important parameter that controls N20 levels, high temperature leading to reduced
N20 emissions. As a result, N20 emission is low (< 20 ppm) in gas- and oil-fired boilers, pulverisedcoal burners and most conventional combustors, whereas fluidised-bed combustion poses a threat
of increased N20 emissions (20-250 ppm). Further augmentation of N20 presence in the atmosphere may result from the use of catalytic convertors in cars and from some NOx control
technologies (e.g. selective non-catalytic reduction). Limestone addition tends to reduce N20
levels and to increase NO emission, whereas increased SO2 levels in the combustor have the
opposite effect. The nature of these interactions is discussed. Due to different design, different
N20/NO formation and destruction pathways may be important in bubbling and circulating FBC's.
In view of uncertainties surrounding the issue of global warming, cautious but prompt N20 control
measures are advocated: (a) improvements in the existing plant (operating conditions, process
control, etc.); (b) research directed to understand N20/NO chemistry in an FBC which would lead
to innovative design of new combustors and their operating r6gimes. Staged combustion, gas
reburning and catalytic enhancement of N20 decomposition seem to be attractive options in N20
control. In view of the existing interactions among individual pollutants and pollution control
measures, an integrated approach to SOx/NOx/N20 abatement is emphasised.
1. INTRODUCTION
T h e m a i n o b j e c t i v e o f t h i s r e v i e w p a p e r is to e x a m i n e f o s s i l f u e l c o m b u s t i o n
as a s o u r c e o f n i t r o u s o x i d e ( N 2 0 ) e m i s s i o n i n t h e c o n t e x t o f e n v i r o n m e n t a l
c o n c e r n s . A n o t h e r g o a l is t o p r o v i d e a n o v e r v i e w o f f u n d a m e n t a l a n d a p p l i c a t i o n - o r i e n t e d r e s e a r c h r e l e v a n t to N 2 0 f o r m a t i o n a n d i t s c o n t r o l .
I n S e c t i o n 2, t h e c o n t r i b u t i o n o f N 2 0 t o b o t h t h e g r e e n h o u s e e f f e c t a n d o z o n e
l a y e r d e p l e t i o n is d i s c u s s e d . N i t r o u s o x i d e is c o m p a r e d w i t h o t h e r g r e e n h o u s e
g a s e s a n d a g l o b a l m a s s b a l a n c e o f t h i s s p e c i e s i n t h e a t m o s p h e r e is p r e s e n t e d .
Analysis of available emission inventories indicates that fluidised-bed combustion (FBC) m a y b e c o m e a m a j o r source of a n t h r o p o g e n i c N20. In the a d v e n t of
F B C c o m m e r c i a l i s a t i o n , t h i s t e c h n i q u e is s u b j e c t e d to t h o r o u g h s c r u t i n y f r o m
t h e p o i n t o f v i e w o f its e n v i r o n m e n t a l i m p a c t . F o r m a t i o n a n d d e s t r u c t i o n o f
N 2 0 in fluidised-bed c o m b u s t i o n are a c e n t r e p i e c e of this paper.
I n S e c t i o n 3, p h y s i c a l a n d c h e m i c a l p r o c e s s e s r e l e v a n t to N 2 0 f o r m a t i o n a n d
d e s t r u c t i o n d u r i n g c o a l c o m b u s t i o n a r e d e s c r i b e d . I t is p o i n t e d o u t t h a t N 2 0

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

emissions as well as abatement policies must be considered in conjunction with


NO,,. A general scheme for the fate of fuel-bound nitrogen is outlined and NzO
chemistry of coal combustion is systematically elaborated upon. Processes
involved in devolatilisation of coal-bound nitrogen are described, and this is
followed by discussion of gas-phase and heterogeneous N20/NO formation and
destruction chemistry. Implications of gas-phase reaction equilibria are also
included.
Against this background, the effects of temperature, fuel type, excess air and
other parameters on NzO/NO emissions from fluidised-bed combustors are
discussed in Section 4. The role of char and limestone is detailed as well as
differences between bubbling and circulating fluidised-bed combustors. A link
is also made between N:O emission and existing NOx and SO~ control measures, with emphasis on their mutual interactions. This leads to the discussion
of comprehensive control strategies and the future of coal firing as a clean
technology for power generation.

2. N20 AND THE ENVIRONMENT


2.1 G r e e n h o u s e effect a n d ozone layer depletion

Nitrous oxide is known to be involved in both the greenhouse effect and in


the destruction of Earth's protective ozone layer. The nat ure of these phenomena is represented schematically in Fig. 1. Generally speaking, the energy

R RADIA
FI(~

UV

PROTECTIVELAYER
-

30 km

o ..... o o ,

02+0""'> ~ + %

12 km

EARTII

Fig. 1. The role played by N:O in the greenhouse effect and in ozone layer depletion.

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

b u d g e t of o u r globe is g o v e r n e d by the b a l a n c e b e t w e e n the a m o u n t of e n e r g y


r e c e i v e d from the sun and h e a t losses to o u t e r space in the form of i n f r a r e d
r a d i a t i o n from the surface. The g r e e n h o u s e effect consists in a b s o r p t i o n of p a r t
of this r a d i a t i o n by some gases p r e s e n t in the a t m o s p h e r e and s u b s e q u e n t
r e - r a d i a t i o n of h e a t b a c k to the Earth. This in t u r n is r e s p o n s i b l e for an
i n c r e a s e in the a v e r a g e t e m p e r a t u r e on the E a r t h , also k n o w n as global
w a r m i n g . The g r e e n h o u s e effect t a k e s place in the t r o p o s p h e r e and the m a i n
g r e e n h o u s e gases are CO2, CH4, N20, c h l o r o f l u o r o c a r b o n s (CFC's), ozone and
w a t e r v a p o u r . T h e a m o u n t of e n e r g y flux re-emitted to the E a r t h is t e r m e d
r a d i a t i v e forcing a n d is u s u a l l y expressed in units of e n e r g y per u n i t of t i m e per
u n i t of a r e a (e.g. in W/m2). It should be n o t e d t h a t the g r e e n h o u s e effect is
a p e r f e c t l y n a t u r a l p h e n o m e n o n w i t h o u t w h i c h life on the E a r t h would be
impossible, or in a n y case v e r y different from w h a t we k n o w it to be now.
W i t h o u t a t h e r m a l b l a n k e t of g r e e n h o u s e gases, the a v e r a g e t e m p e r a t u r e on
o u r p l a n e t would be - 18~C i n s t e a d of the o b s e r v e d + 15C [1]. T h u s it is not the
g r e e n h o u s e effect itself t h a t raises e n v i r o n m e n t a l c o n c e r n s but its e n h a n c e m e n t due to h u m a n a c t i v i t i e s w h i c h b r i n g a b o u t an i m b a l a n c e w i t h i n the
delicate s y s t e m of e n e r g y sources and sinks.
D e p l e t i o n of the s t r a t o s p h e r i c ozone l a y e r is likewise the r e s u l t of a maninduced i m b a l a n c e a m o n g the sources a n d sinks of ozone. S t r a t o s p h e r i c ozone
c o n s t i t u t e s a n a t u r a l shield p r o t e c t i n g us from excess u l t r a v i o l e t (UV) radiation. U V r a y s are believed to c a u s e an i n c r e a s e d risk of skin c a n c e r and
possibly also s u p p r e s s i o n of the h u m a n i m m u n e s y s t e m [2]. R e m o v a l of ozone
from the s t r a t o s p h e r e m a y also affect the d i s t r i b u t i o n of s o l a r h e a t i n g and thus
c o n t r i b u t e to global w a r m i n g . The n a t u r a l cycle of ozone f o r m a t i o n and
d e s t r u c t i o n has been k n o w n a b o u t since the 1930's and it c a n be r e p r e s e n t e d by
the following set of reactions:
UV

Formation:

Destruction:

02

, O+O

(1)

O + 02 -

, 03

(2)

03 + O

, 20 2

(3)

Ozone loss ( r e a c t i o n 3) m a y be c a t a l y s e d by a n u m b e r of t r a c e species, the


c o n c e n t r a t i o n of w h i c h in the s t r a t o s p h e r e h a s s u b s t a n t i a l l y i n c r e a s e d due to
h u m a n input. S u c h a c a t a l y t i c cycle c a n be r e p r e s e n t e d s y m b o l i c a l l y as follows,

03+*

, *0+02

(4)

*0+0

, *+02

(5)

w h e r e * stands for an a c t i v e c o n s t i t u e n t , e.g. a c h l o r i n e a t o m (C1) or a m o l e c u l e


of nitric oxide (NO).
As i n d i c a t e d in Fig. 1, n i t r o u s oxide is a s o u r c e of NO m o l e c u l e s in the
s t r a t o s p h e r e w h e r e t h e y play a role in the d e s t r u c t i o n of ozone. N i t r i c oxide

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

would otherwise be unlikely to reach such altitudes because of its


short atmospheric lifetime (0.5-5 days [3] as opposed to 150-160 years for N20
[1, 3]). Thus the contribution of N20 to ozone layer depletion is indirect but
quite potent. Almost half of the ozone loss in the stratosphere is attributed
to NOx [4] and, since N20 is its major source, a doubling in the atmospheric
N20 concent r a t i on could result in a 12% decrease in total stratospheric
ozone [5].
E n h a n c e m e n t of the greenhouse effect due to human activities is difficult to
quantify. Uncertainties are manifold. They involve difficulties in reliable
measurement and averaging of temperature and concentrations within a dynamically changing atmosphere. Regional and seasonal changes need to be
accounted for and very few reliable data are available from measurements
performed in the previous century. Time scales of events are enormous--centuries and m i l l eni a- - and it has been argued t hat the observed trends are nothing
but normal fluctuations resulting for instance from variations in solar activity
or from volcanic eruptions (e.g. [6-9]). Nevertheless, progress has been made
and fairly reliable data are now available on the extent of global warming and
on changes in concentrations of greenhouse gases. It is believed t hat a real
warming up of the globe of 0.3 0.6 K has taken place over the last century [1].
Analysis of air trapped in Antarctic ice cores at various depths made it possible
to determine the evolution of atmospheric composition over more than 2000
years [10-14]. The results of such measurements are presented in Fig. 2, from
which it is evident that prior to industrialisation, global concentrations of N20
were relatively constant at around 285 ppb. Following industrialisation, N20
levels have been increasing steadily at about 0.3% per year, to reach the

350

.1

325

300
o

.. s

275

:,

a.

..

.".'l.'~,',,.,"
ol

250

l
500

I
1000

Date of saml)le

I
1500

I
2000

lyear AI)I

Fig. 2. N20 measurements from ice-core samples (after IPCC, 1990 [1]; data by Khalil and
Rasmussen [10], Etheridge et al. [13] and Zardini et al. [14]).

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

current value of approximately 310 ppb (N20 is the most abundant atmospheric
nitrogen species after molecular nitrogen). More direct measurements performed in the 1970's and 1980's confirmed this trend [15] and it is fair to assume
that this change in N20 concentration is due to human activities.
The impact of both global warming and ozone destruction on Earth's climate
and human life is even more difficult to assess. Predictions are made primarily
by means of computer-aided modelling of climate, and this approach has been
used with varying success in the past. The first, one-dimensional, models
turned out to be too crude to predict the dramatic changes in the ozone layer
that occurred in the 1980's (the so-called "ozone hole"). The computer models
were later expanded to three dimensions and improved to include phenomena
originally overlooked, but they still contain inherent uncertainties arising
both from incomplete understanding of complex atmospheric chemistry and
from lack of knowledge of future emissions of trace species [16]. Generally
speaking, global warming is expected to affect sea level, precipitation, soil
moisture, vegetation, ocean circulation, wind patterns, frequency and distribution of storms, floods, etc. Of course the influence of global warming will be
different at different places on the Earth. Rising sea level may cause a lot of
concern in lowland countries, such as The Netherlands, but it is conceivable
that a warmer climate would improve agriculture and reduce energy consumption in the United Kingdom, Scandinavia, Russia or China. It is often argued,
however, that no disturbance in the ecosystem can be tolerated because of the
delicate nature of the environment. Demands are put forward that legislation
should be passed to curb emissions of greenhouse gases, notably of CO2. This in
turn raises a lot of opposition among companies involved in energy generation.
Conflicting interests make it difficult to reach an international consensus,
which adds a political dimension to the already controversial issue.
Climatic response to both the greenhouse effect and stratospheric ozone
depletion is further complicated by a number of mutual interactions between
these phenomena as well as by some feedback mechanisms. The following
example of such an interaction and feedback can be given: N20 destruction by
photolysis is enhanced by increased UV radiation [3]. In this way, a result of
ozone layer depletion mitigates itself through negative feedback and, additionally, reduces the greenhouse effect. Another interaction involves cooling of the
stratosphere which results from tropospheric heat entrapment by greenhouse
gases. Cooling of the stratosphere slows down the kinetics of ozone destruction
[16]. As mentioned before, stratospheric ozone depletion also brings about an
increase in the flux of solar energy to the Earth, in this way augmenting the
warming trend. Yet a not he r feedback occurs when the atmosphere warms,
causing increased water evaporation. Water vapour is an important greenhouse gas and will thus amplify the warming. The production of nitrous oxide
in soil via denitrification and nitrification also increases with soil temperature
[17]. Moreover, global warming is also likely to result in a higher occurrence of
drought, increased frequency of burning, which in turn leads to larger N20
emissions [1]. Other feedbacks occur via interactions with snow, sea ice, clouds
and the biosphere. In general, negative feedbacks can reduce the warming but
cannot produce a global cooling [1].

M.A. Wdjtowicz et al./Fuel Processmg Technol. 34 (1993) 1 71

2.2 C o n t r i b u t i o n o f N 2 0 to the g r e e n h o u s e effect

The r a d i a t i v e forcing of the individual g r e e n h o u s e gases can be c a l c u l a t e d


with m u c h more confidence t h a n the climatic c h a n g e it causes because the
f o r m e r avoids the need to e v a l u a t e a n u m b e r of poorly u n d e r s t o o d a t m o s p h e r i c
responses. The following factors d e t e r m i n e the ability of a given species to
exhibit the g r e e n h o u s e effect:
(a) the i n f r a r e d a b s o r p t i o n s t r e n g t h and the r a n g e of w a v e l e n g t h s in which IR
a b s o r p t i o n occurs
(b) o v e r l a p p i n g of a b s o r p t i o n bands with those of o t h e r g r e e n h o u s e gases
(c) the q u a n t i t y of gas in the a t m o s p h e r e
(d) gas lifetime in the a t m o s p h e r e
(e) the period of time over which the r a d i a t i v e forcing is considered
A few comments are due on the above list. First of all, different gases exhibit
different r e l a t i v e g r e e n h o u s e p o t e n t i a l s on a mol-per-mol (or kg-per-kg) basis.
In this way, the most a b u n d a n t c o n s t i t u e n t is not necessarily the most p o t e n t
g r e e n h o u s e c o n t r i b u t o r . Secondly, since the r a d i a t i v e forcing is g e n e r a l l y
non-linear in absorber amount, the forcing should be r e l a t e d to c o n c e n t r a t i o n
changes r a t h e r t h a n to c o n c e n t r a t i o n itself, preferably in the r a n g e of c u r r e n t
c o n c e n t r a t i o n of a given species. Of course, assessment of the p o t e n t i a l impact
of f u t u r e emission levels is even more i m p o r t a n t and thus a t m o s p h e r i c lifetime
needs to be considered. For example, a short-lived gas which has a strong (on
a kg-per-kg basis) g r e e n h o u s e effect m a y in the short term be more effective at
c h a n g i n g the r a d i a t i v e forcing t h a n a w e a k e r but longer-lived gas. Over longer
periods, however, the i n t e g r a t e d effect of the w e a k e r gas m a y be more significant due to its persistence in the a t m o s p h e r e [1].
Despite a p p a r e n t complexity, a n u m b e r of progressively more a d e q u a t e
indices h a v e been developed as a measure of the r a d i a t i v e forcing. It has
become c o m m o n p r a c t i c e to r e l a t e the s t r e n g t h of the r a d i a t i v e forcing to t h a t
of CO2. The most a d v a n c e d index to date seems to be the Global W a r m i n g
P o t e n t i a l ( G W P ) [1], defined as follows:
If

aicidt
GWP

~ o

(6)

aco~ Cco~dt
0

In this expression, ai is the i n s t a n t a n e o u s r a d i a t i v e forcing due to a unit


i n c r e a s e in the c o n c e n t r a t i o n of gas i; ci is c o n c e n t r a t i o n of gas i r e m a i n i n g at
time t after its release; and tf is the time (number of years) over which the
c a l c u l a t i o n is performed. Global W a r m i n g P o t e n t i a l defines the time-integrated w a r m i n g effect due to an i n s t a n t a n e o u s release of unit mass of a given
gas in t o d a y ' s atmosphere, r e l a t i v e to t h a t of CO2. Thus the c o n t r i b u t i o n to
global w a r m i n g over a period of time considered can be e v a l u a t e d as the
p r o d u c t of G W P and the a m o u n t of g r e e n h o u s e gas emitted [1].

M.A. WSjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

r~

rd

~?.

,,,,,~

e'~

~8

8
o

8~

'~
e~

~.~_

.~

".~

"''~

~,~

~-~o
b-,

b-

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

Characteristics related to the radiative forcing for several gases are presented in Table 1. The data show that CO2 is the least effective greenhouse gas
per kilogramme emitted but its contribution to global warming is most significant due to large amounts released in the atmosphere. It is also evident t hat
the warming potential of CFC's is enormous and that, quite fortunately, it is
mitigated by low atmospheric concentrations of these species. Nitrous oxide is
clearly a strong greenhouse gas and exceptionally long-lived in the atmosphere. Any reductions in its amount now will not bring about any improvement until after many decades. Nitrous oxide's persistence in the atmosphere is
illustrated in Fig. 3. Its very long lifetime and strong radiative forcing are what
raises environmental concerns about NeO.
Generally speaking, control over the enhanced greenhouse effect cannot be
accomplished by a reduction in the emission of a single gas. Contributions are
well-distributed, with the trace greenhouse gases accounting for about as
much radiative forcing as CO2 alone. Therefore, a concerted effort is needed to
cut down emissions of all the species involved. The same applies to the
depletion of stratospheric ozone [3]. It has been estimated t hat in order to
stabilise atmospheric concentration of N20 at the current level, a 70-80%
reduction in human-made emissions of this gas is required [1].

2.3 Inventory of N 2 0 emissions

Table 1 shows that the mean atmospheric concentration of N20 is currently


about 310 ppb, which corresponds to a reservoir of approximately 1500 Mt(N).
The annual accumulation of N20, corresponding to a 0.2-0.3% increase in its
concentration, is about 3-4.5 Mt(N). The major atmospheric sink for N20 is
photochemical decomposition in the stratosphere, tropospheric sinks (aquatic
and soil systems) are believed to be insignificant [20, 21]. Photolysis of N20

1.0

I'420

0.5

crc-,,

50
Time

1130

150

[years]

Fig. 3. Fraction of emissions of different greenhouse gases remaining in the atmosphere as


a function of time (Rodhe [19] (('~1990 by the AAAS).

10

M.A. WOjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

occurs a c c o r d i n g to the following reactions:


N20 + hv
N20+O

--

, N2 + O

(7)

, N2+O2

(8)

The total r a t e of decomposition is estimated to be about 7 13 M t ( N ) / y e a r [1]


and a simple mass balance shows t h a t the sources of N20 a m o u n t to a b o u t
10-17.5 Mt(N)/year. This r e p r e s e n t s a 30% excess of sources over sinks and
h u m a n activities are believed to be the m a j o r cause of this imbalance.
The global N 2 0 i n v e n t o r y is poorly quantified. A p a r t from difficulties inh e r e n t in d e t e r m i n a t i o n of global fluxes of N20 from such sources as oceans,
soils, or from biomass burning, m a n y d a t a led to spurious results due to
a sampling artifact t h a t was not realised until 1988 [22]. The artifact involved
f o r m a t i o n of N20 from NO and SO2 in sampling flasks before analysis was
performed. This effect is now well u n d e r s t o o d and improved p r o c e d u r e s for
N20 sampling and analysis have been developed and validated. It is i n t e r e s t i n g
to note, however, t h a t no c o n v e n i e n t on-line m e t h o d for N20 m e a s u r e m e n t is
k n o w n to date, which makes it difficult to m o n i t o r N 2 0 emissions in an
i n d u s t r i a l e n v i r o n m e n t . Most m e a s u r e m e n t s are still performed using cumbersome grab sampling, followed by gas c h r o m a t o g r a p h i c analysis. Infrared and
p h o t o a c o u s t i c a n a l y s e r s are available but in these, c o r r e c t i o n s need to be made
for i n t e r f e r e n c e from o t h e r gases present in the m i x t u r e (notably COz and
H20). Gas p r e t r e a t m e n t , such as drying or trapping i n t e r f e r i n g species, is also
used as an alternative.
Sources of a t m o s p h e r i c N20 can be divided into n a t u r a l and a n t h r o p o g e n i c .
The former involve emission from soils, oceans and f r e s h w a t e r as well as r a t h e r
small a m o u n t s of N 2 0 g e n e r a t e d by lightning. It is estimated t h a t n a t u r a l
sources a c c o u n t for about 60-70% of total N20 emission into the a t m o s p h e r e
[23, 24].
One c o n t r i b u t i o n to the a n t h r o p o g e n i c sources of N 2 0 comes from b u r n i n g
fossil fuels, e i t h e r in oil-, gas- or coal-fired power stations ( s t a t i o n a r y sources),
or in the i n t e r n a l c o m b u s t i o n engines of cars, lorries, etc. (mobile sources). The
largest a m o u n t s of N 2 0 released into the a t m o s p h e r e due to h u m a n activities
come from land c u l t i v a t i o n and are associated m a i n l y with the use of fertilisers. Biomass b u r n i n g and the m a n u f a c t u r e of adipic acid (used in n y l o n
production) complete the i n v e n t o r y of k n o w n sources.
Detailed description of all the above-mentioned N : O sinks and sources is
beyond the scope of this paper. For more e x h a u s t i v e discussion, the r e a d e r is
referred to the cited literature, n o t a b l y to the IPCC r e p o r t [1] and to the work
by Sloss [3] and Levine [18]. In this study, we shall limit ourselves to mere
compilation of the k n o w n sources and t h e i r strengths. Even this will be done
only to the e x t e n t n e c e s s a r y to e v a l u a t e the relative c o n t r i b u t i o n of coal
c o m b u s t i o n to the total emission of N20.
Global emissions of N : O are shown in Fig. 4. As m e n t i o n e d before, d a t a on
the s t r e n g t h of N:O sources t h a t are available in the l i t e r a t u r e v a r y widely,
and we have chosen to reflect this v a r i a t i o n in our compilation. In this way,
ranges of p a r t i c u l a r emissions are p r e s e n t e d r a t h e r t h a n m e a n values, and it

11

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


7

ANTHROPOGENIC

6-5

5~

7-

NATURAL

100 % FBC

5
I
I

~, 3

+
2
z ~

i
[

t
I
I
I
I
I
I
I
I

[]

2.6

ill J
1.4

0.85

00.666

040

lowerestimate I
higher estimate I

fossil fuels fossil fuels adipic acid


(stationary) (mobile)

1.0

biomass
land
burning cultivation

soils

oceans & lightning


freshwater

Fig. 4. Global sources of N20 (after [1, 23 37]). 100% FBC is a prediction for a hypothetical
situation in which all the coal and peat is combusted in FBC's. In estimating emissions from
mobile sources burning fossil fuels (cars, lorries, etc.), it was assumed that 50% of the petrol
engines were equipped with medium-aged catalytic convertors (De Soete, 1990 [23]).

s h o u l d be n o t e d t h a t only d a t a free of the N 2 0 s a m p l i n g a r t i f a c t h a v e b e e n


included. In a few cases, emission e s t i m a t e s r e p o r t e d were deemed excessively
large. If no c o n v i n c i n g s u b s t a n t i a t i o n was p r o v i d e d in t h e i r support, t h e y h a v e
b e e n disregarded.
It c a n be i n f e r r e d f r o m Fig. 4 t h a t the t o t a l a m o u n t of N 2 0 k n o w n to be
e m i t t e d into the a t m o s p h e r e is 5.3--14.4 M t ( N ) / y e a r . C o m p a r i s o n with the
e s t i m a t e derived f r o m the global b a l a n c e of sinks, sources a n d a t m o s p h e r i c
a c c u m u l a t i o n (10-17.5 M t ( N ) / y e a r ) i n d i c a t e s the existence of some deficit in
the N 2 0 budget. (The deficit was e v e n l a r g e r before the c o n t r i b u t i o n s from
mobile c o m b u s t i o n s o u r c e s and adipic acid m a n u f a c t u r e were included.) W h e n
first realised, this d i s c r e p a n c y b e t w e e n the s t r e n g t h s of k n o w n and e s t i m a t e d
NzO s o u r c e s s t i m u l a t e d a lot of r e s e a r c h leading to i d e n t i f y i n g new and b e t t e r
q u a n t i f i c a t i o n of old sources. In this way, N 2 0 h a s r e c e n t l y b e e n found to be
a b y - p r o d u c t of n y l o n m a n u f a c t u r i n g , m o r e specifically of the adipic acid
p r e p a r a t i o n [25]. In this process, adipic acid is formed by o x i d a t i o n of cycl o h e x a n o l c y c l o h e x a n o n e m i x t u r e s w i t h nitric acid, w i t h N 2 0 b e i n g one of the
r e a c t i o n products. It is e x p e c t e d t h a t m o r e s u c h discoveries will e m e r g e in the
future, leading to a closure in N 2 0 m a s s b a l a n c e .
In spite of u n c e r t a i n t i e s in the N 2 0 i n v e n t o r y , Fig. 4 shows t h a t p o w e r
g e n e r a t i o n from fossil fuels is c e r t a i n l y not the m a i n p e r p e t r a t o r of the enh a n c e d g r e e n h o u s e effect; at least not at the m o m e n t . In the future, h o w e v e r ,
w h e n fluidised-bed c o m b u s t i o n b e c o m e s a m a j o r c o a l - b u r n i n g t e c h n o l o g y ,
N 2 0 emissions m a y b e c o m e t r u l y d r a m a t i c . A c c o r d i n g l y to De Soete [23],
the c o n t r i b u t i o n f r o m fossil fuel c o m b u s t i o n would i n c r e a s e to a b o u t

12

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

4.2-5.2 Mt(N)/year if all the coal and peat combusted nowadays were burnt in
FBC's. It is evident from Fig. 4 that such a change would make combustion of
fossil fuels the strongest source of N20, and would more than double anthropogenic contributions to global N20 input. Although commercialisation of
FBC technology is likely to be gradual, concern about increased N20 emissions remains valid. Even at cur r ent emission levels, Fig. 4 shows that except
for N20 release related to land cultivation, which is probably somewhat
higher, all other anthropogenic sources are of comparable strength and thus
should all be controlled.
Finally, the question of N20 emissions from various types of combustors
needs to be addressed. Extensive compilation of such data for more than 150
cases has recently been performed by Hulgaard [38]. He concludes that, in
general, N20 emission is low for natural gas-fired boilers and turbines
(< 2 ppm), utility boilers fired with oil (0-5 ppm), as well as for conventional
pulverized-coal burners (< 5 ppm). Emission levels from grates are somewhat
higher ( < 20 ppm), whereas flue gas from FBC's (both bubbling and circulating)
contains between 20 and 250 ppm N20. Similar figures are also quoted elsewhere [23, 39]. Circulating and bubbling FBC's have approximately the same
level of N20 emissions [40]. The data clearly show t hat N20 production in
FBC's is at least an order of magnitude higher than that coming out of
conventional boilers. This is offset by improved control of SOx and NOx
emissions from fluidised beds, which is one of the main reasons why FBC
technology has been developed.
In summary, the following statements can be made with reasonable confidence:
(1) Atmospheric concentration of greenhouse gases has been increasing
steadily for many decades and this increase is due to human activities.
(2) A slight increase ( ~ 0.3 0.6 K) in the mean temperature of the Earth has
occurred over the last century.
(3) N20 is a strong, long-lived absorber of infrared radiation and as such
contributes to enhancement of the greenhouse effect.
(4) N20 acts as a nitric oxide precursor in the stratosphere and in this way
plays an important role in ozone layer depletion.
(5) The contribution of N20 to the greenhouse effect has been rising steadily
and was about 6% in the 1980's.
(6) N20 raises environmental concerns due to (a) its long atmospheric lifetime,
and (b) the possibility of a massive increase in its emission upon switching
to fluidised-bed combustion.
The following issues need further clarification:
(1) It is not clear whether the global warming observed over the last century is
due to human activities, in particular due to the enhanced greenhouse
effect.
(2) Environmental response to global warming is unknown and it is still
debated whether or not, and to what extent, warming of the climate should
be resisted.
In view of the above, a prudent, well-balanced approach to the N20 question
is needed. As it was rightly pointed out by Davis [41], we are dealing with

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

13

a three-dimensional problem involving energy, environment and economics.


Firstly, the world needs energy, and the demand for this commodity is bound to
be continually growing. Secondly, the environment has to be protected and,
finally, both needs ought to be satisfied at the lowest possible cost. We believe
th at given the extremely slow response of N20 atmospheric levels to changes
in its emission, doing nothing and waiting for a better understanding of
atmospheric chemistry of this species may be risky. It is interesting to note t hat
legislation limiting the use of chlorofluorocarbons, the so-called Montreal
Protocol, was agreed upon about a fortnight before the first unequivocal proof
of CFC's harmful effect on the ozone layer was furnished [16]. Now it is
well-known that even further reductions in CFC emissions are necessary. Back
in 1987, however, or even earlier in 1985 when the Vienna Convention was
signed by 20 nations, scientific concern about stratospheric ozone arose entirely from theory and from (often badly inaccurate) modelling of the underlying
atmospheric processes. Should this then serve as a lesson for the greenhouse
effect?
Referring to N20 emissions from coal combustion, substantial reduction in
N20 levels can be accomplished by improving technologies. Fluidised-bed
combustion should not be dreaded; instead, modifications in the design and
operation of FBC's should be introduced in such a way so that N20 abatement
is attained without sacrificing low NOx and SO~ levels. To this end, the
knowledge of N20 formation and destruction under fluidised-bed conditions
should be advanced and applied as far as possible.

3. N20 CHEMISTRY IN COMBUSTION OF COAL


The chemistry of nitrous oxide formation and destruction during coal combustion is complex and still insufficiently understood although great strides
towards improving this situation have been made in recent years [42-49].
Emissions of N20 should be considered in the broad context of the fate of
coal-bound nitrogen (coal-N) during the combustion process. It seems most
appropriate to study N20 formation and destruction in close relation to mechanisms responsible for the release of other oxides of nitrogen as well as
molecular nitrogen. Failure to do so may not only result in fragmentary
knowledge of nitrogen evolution during combustion, but may also lead to
adopting pollution-control measures that reduce emissions of one pollutant at
the expense of increased levels of others. In fact, fluidised-bed combustion
serves as a good example of such a one-sided approach. At the time when the
enhanced greenhouse effect and stratospheric ozone were not an issue,
fluidised-bed combustion emerged as the result of a successful approach to the
problem of NOx/SOx abatement. Not recognised as an environmental pollutant,
N20 was for a long time ignored and its increased levels in FBC's unnoticed.
A general overview of the N20/NO reaction scheme is presented in Fig. 5,
with major reactions listed in Tables 2 and 3. It should be pointed out that
although this picture may not be complete, it is complex enough to justify some
simplification. Not all the reactions shown are of equal importance and the

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

14

NOIN=OFORMATION

Di~WOLATIL|SATION

NO/N~O

IICN, NIl s
N,

i (volatile-N)

NO/NzODES'rRUCI'ION

FINALEMISSION
NO/N~

gas-phase reduction

_N~
(also

tar-Nt

t '
X
I

CATALYST

helerogeneous renctlons

_N~O

t
CATALYS
f/

COALPARTICLE
Fig. 5. Formation and fate of NO and N20 during combustion of coal.
role they play in the overall N20/NO formation mechanism varies depending
on whether coal flames, fluidised-bed combustors, or entrained-flow reactors
are considered.
As discussed above, fluidised-bed combustion is the only technique involved
in electric power generation which, despite its unquestionable advantages, has
the potential to cause a serious imbalance in the N20 budget. This is why the
analysis of N20 release carried out below is focused on FBC, although the
described processes are certainly relevant to other methods of burning carbonaceous materials.
It is fair to assume that in fluidised-bed combustion, the only contribution to
final NOx/N20 emissions comes from fuel-bound nitrogen. This assumption is
valid because of the relatively low combustion temperature prevailing in
FBC's (typically about 1100 K), which prevents formation of nitrogen oxides
from oxygen and nitrogen present in the gas mixture (the so-called "thermal
NO") [71 74]. Nitric oxide makes up 90-95% of the NOx emissions from fossil
fuel combustion [3]. Under normal operating conditions, very little NO2 is
detected in FBC's and NOx consists almost exclusively of NO [75].
Referring to Fig. 5, coal-N transformation during combustion can be divided
into three stages: coal devolatilisation and pyrolysis, N20/NO formation (both
in the gas phase and at the surface of char), followed by N20/NO destruction
(mainly reduction to N2). The latter process can again be either homo- or
heterogeneous. The amount of nitrogen oxides that survives the last stage is
carried out of the system with flue gases. In this way, a four-component scheme
is suggested as a framework for the analysis: (a) gas-phase N20/NO formation,

M.A. W6jtowicz et al,/Fuel Processing Technol. 34 (1993) 1- 71

15

TABLE 2
M a i n r e a c t i o n s i n v o l v e d in t h e N 2 0 / N O c h e m i c a l scheme (after [38, 43, 44, 48 64]). Boldface
p r i n t e d r e a c t i o n s r e p r e s e n t m a j o r r e a c t i o n p a t h w a y s ; t h e r a t e expression in t h e A r r h e n i u s
form: k = A T~exp( - E/RT)
#

Reaction

A
(cm 3/(tool s))

fi
(-)

E
(kJ/mol)

Ref.

1. G a s - p h a s e N 2 0 / N O f o r m a t i o n
(a) N 2 0 / N O from NH3
R1
NH3 + OH -* NH2 + H 2 0
R2
NH, +OH-~ NH+H20
R3
NH2+H-* NH+H2

4.7 E + 0 6
4.0 E + 06
4.0 E + 13

1.90
2.00
0.00

2.08
4.16
30.35

[57]
[56]
[58]

R4
R5
R6

NH+H-~ N+H2
N+OH-~ NO+H
NH+O-~NO+H

3.2 E + 13
3.8 E + 13
7.8 E + 13

0.00
0.00
0.00

1.36
0.00
0.00

[59]
[56]
[60]

R7
R8

NH2 + NO -* NzO + H2
NH + N O -* NzO + H

5.0 E + 13
4.3 E + 1 4

0.00
-0.50

103
0.00

[63]
[50]

R9
R10
Rll
R12
R13
R14

(b) N 2 0 from HCN


H C N + O -* NCO + H
HCN+OH-~ CN+H20
C N + O 2 -* N C O + O
HCN+OH-* HNCO+H
H N C O + O H -* N C O + H 2 0
NCO+NO-~ N20+CO

1.4
3.3
1.0
2.0
2.6
1.0

E + 04
E + 12
E + 13
E-03
E + 12
E + 13

2.64
0.00
0.00
4.00
0.00
0.00

20.9
45.6
0.00
4.16
23.3
- 1.66

[56]
[38]
[61]
[56]
[62]
[56]

R15
R16

(c) NO from HCN


NCO+O~NO+CO
NCO+OH---+ N O + C O + H

2.0 E + 13
1.0 E + 13

0.00
0.00

0,00
0.00

[56]
[56]

R17
R18

(d) T h e r m a l NO
N2+O~NO+N
N+O2--* N O + O

6.4 E + 09

26.27

[56]

2. G a s - h a s e NzO/NO d e s t r u c t i o n
(a) Direct
R19 N 2 0 + H - , N 2 -b OH
R20 NzO + OH ~ N 2 n o 2
R21 N z O + M - - * N 2 + O + M
R22 N 2 0 + CO --* N 2 + CO2
R23
R24

R25
R26
R27

N H 2+NO--* N z + H 2 0
N+NO-* N2+O
(see also R7, R8 a n d R14 above)

(b) I n d i r e c t
NCO+H--* NH+CO
N C O + H 2 0 -+ H N C O + O H
NCO + H 2 --~ HNCO + H

7.6
2.0
1.6
2.5

E+
E+
E+
E+

13
12
14
14

See R24
1.00

0.00
0.00
0.00
0.00

63.6
41.6
215
192

[56]
[56]
[56]
[64]

6.2 E + 15
3.3 E + 12

- 1.25
0.30

0.00
0.00

[56]
[56]

5.0 E + 13

0.00
See R13
0.00

0.00

[561

8.6 E + 12

3. G a s - p h a s e r e a c t i o n s i n v o l v i n g N H 3- a n d H C N - b a s e d s p e c i e s
R28 H C N + O ~ N H + C O
3.5 E + 0 3
2.64
R29 H N C O + H - - * N H 2 + C O
1.1 E + 1 4
0.00
(See also R 1 - R 4 a n d R25
above)

37.4

[56]

20.8
53.2

[561
[601

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

16

gS.=.~

,,.,t

oo

%1

. ~

r'--

0
~ .~
,-~
~ 2.~
M

~~

..,

.~ ~ "

.~"~

--

V_,

0
Z
['-

"~ "i Z
m
e-,

.,-i
m

..,
0

e3

<

M.A.

W6jtowicz

et a l . / F u e l

Processing

Technol.

17

34 (1993) 1 - 7 1

,o

oq.
0

~.

.~.
,,6

o'x

~.
oo

,.~
c5

eq

+
o.

IZ

,.q.

r.q

~-

.) G)

<~
Oo
oz

(.9
~
~

oz

k~
.

kaz.

dz~

dz~

o-i
G

o~
~

=
+

oeq

+
-c?~

c?
~

o~a+

"#

=
+

+
+

+++~

o~

ZZZ~.~

~+o

o_
Z

,T

Q
Z

T'd

Z
eq

M.A. WSjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

18

e-

,m

i
0

e.
0

e-

0
l

"Fs
+

:f

"4

e"

~z

-~

z~

*d

itzt

0
tt

e~

e"

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

19

(b) gas-phase N20/NO destruction, (c) heterogeneous N20/NO formation, and


(d) heterogeneous N20/NO destruction. Devolatilisation can be viewed as
a physicochemical process governing the relative amounts of original coal-N
that will react along the gas-phase versus heterogeneous routes.
In addition, there are a number of reactions which directly involve neither
N20 nor NO but which have a strong effect on N20/NO chemistry. They
typically involve gas-phase precursors of N20/NO, such as NH3 and HCN, or
species known to be instrumental in NaO/NO destruction, e.g. CO. Such
reactions can take place in the gas-phase or on solid surfaces and they form
a separate category not shown in Fig. 5.
The main nitrogen-containing products of coal devolatilisation and pyrolysis are hydrogen cyanide (HCN), ammonia (NH3), N2 as well as tar- and
char-bound nitrogen. Hydrogen cyanide and NH3 play an important role
acting as gas-phase precursors for both N20 and NO and these reaction
pathways have been extensively studied (see the discussion below). Reactions
of nitrogenous species can be substantially accelerated, in some cases also
modified, by catalytic effects associated with the presence of solid particles in
the reacting mixture. In the current practice of fluidised-bed combustion, these
are usually ash, limestone or dolomite particles, but in future applications,
other sulphur-capture sorbents or de-NO~/N20 catalysts can be used.
Combustion of char is another source of nitrogen oxides. The relative
contributions of NOx/N20 formed via gas-phase versus heterogeneous routes
depend on the type of fuel combusted, reaction temperature, excess oxygen
level, etc. Apart from NO and N20 resulting from combustion of char, HCN
and NHs may also be formed from char-N, subsequently reacting in the gasphase to yield NO and N20.
Reduction of NO and N20 to molecular nitrogen involves chemical processes
of great practical interest as they constitute a powerful mechanism of
NOdN20 destruction. The reduction may be either heterogeneous, solidcatalysed or occurring on char surface, or it may take place in the gas phase. In
the latter case, H and OH radicals present in the combustion zone are particularly strong reducing agents for N20. As indicated in Fig. 5, incomplete
reduction of NOx may contribute to increased N20 emission levels.

3.1 D e v o l a t i l i s a t i o n a n d p y r o l y s i s

Devolatilisation, which to a large extent is a result of temperature-induced


changes in coal's organic structure (pyrolysis), is a process of paramount
importance for pollution control strategies and combustor design. Devolatilisation kinetics, combined with the rates of char and volatiles combustion, bed
mixing patterns and residence times of the gas and particulate phases, constitute input information for the modelling and design of coal processing plant.
Since pyrolysis is relevant to nearly all coal conversion processes such as
(hydro) gasification, combustion, liquefaction, or carbonisation, it is not surprising that the devolatilisation behaviour of coal has been extensively studied
and a number of review articles are available on this subject [76 82]. In what

20

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

follows, only the evolution of nitrogenous species is addressed, in particular


the distribution of coal-N between the volatiles and char.
The devolatilisation behaviour of a lignite and a bituminous coal was
studied by Pohl and Sarofim [83]. They found that nitrogen-containing species
in coal were relatively stable, in the sense that most of the nitrogen evolution
occurred at temperatures above 1223 K, whereas most of the total weight of
coal was lost below this temperature. Moreover, about 10% of the coal weight
had been lost before nitrogen evolution began. This implies that nitrogen is
present in coal mainly in the form of relatively stable aromatic structures and
that its release is delayed until ring rupture occurs. (In the meantime, other
volatiles are driven off as a result of the breakup of side chains and aliphatic
links.) Destructive pyrolysis and oxidation experiments [84-92] as well as
X-ray photoelectron spectroscopy (XPS) studies of nitrogen functionalities in
coals [93 96] indeed suggest that the primary nitrogen-containing ring structures are those of pyrrolic (five-membered) and pyridinic (six-membered) hature. According to Pohl and Sarofim [83], ring rupture is followed by condensation reactions, leading to the formation of multiple rings with internal carbon
atoms. Nitrogen, on the other hand, is thought to occupy at this stage less
stable positions at peripheral atoms of multiple-ring structures. This description is consistent with the results of experiments involving extensive heattr eatment of coal in a crucible. The data show continuous release of nitrogen
after all other volatiles have already evolved (zero weight loss).
Solomon and Colket performed rapid pyrolysis experiments using a heatedgrid reactor [97]. They found that no substantial amounts of NH3 and HCN
were formed at low temperatures (<973 K) and that tar-N was the main
nitrogenous volatile product under these conditions. Furthermore, nitrogen
concentration in the tar was found to be constant and identical to the concentration of nitrogen in the parent coal over the whole temperature range studied
(i.e. up to 1273 K). It was concluded that the organic structure of tar was
similar to that of the parent coal and that tar-N was evolved in the same form
and concentration as in the parent material. This was later confirmed by
toluene solubility tests performed on tars by Baumann and M611er [98]. They
found that the toluene insoluble fraction (heavy and highly aromatic) included
most of tar-N. It follows then that the low-temperature char must also resemble
the parent coal, which was indeed reflected in the nearly identical nitrogen
contents found in the coal and its low-temperature char [97].
As temperature increases, the nitrogen content in char increases due to the
evolution of non-nitrogenous volatiles. A further increase in the severity of
pyrolysis leads to an eventual decrease in char nitrogen content upon inception of ring rupture. In this way, coal-bound nitrogen evolves initially in heavy
tar molecules without significant change in molecular structure, followed by
secondary nitrogen release due to ring rupture. These processes have distinctly
different kinetics, the reaction rate of the latter being about two orders of
magnitude lower than the rate of initial devolatilisation.
Retention of nitrogen in char was found to decrease with temperature [83].
It is interesting to note that at temperatures characteristic of fluidisedbed combustion (i.e. about 973-1273 K), a substantial proportion of original

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

21

coal-N is retained in the char (typically above 60%). These results were also
confirmed by other studies [97 101]. Nitrogen retention in char seems to be
highest for high-rank coals [98, 102, 103], which can be explained by the high
thermal stability of the condensed multi-ring aromatic structures in high-rank
materials.
A detailed study of the distribution of nitrogen-containing compounds between the char, tar and volatiles was undertaken by Baumann and M611er [98].
In contrast to previous studies, a fiuidised-bed pyrolyser was used in this work.
The results indicate that at relatively low temperatures (573-1173 K, depending on coal type), the N/C ratio is usually higher in chars than in the parent
coals. At higher temperatures (typically above 873-973 K, the exact value
again depending on coal type), the N/C ratio declines, presumably due to ring
rupture. This provides further evidence for the above-described mechanism of
coal-N evolution during pyrolysis.
The distribution of coal-N between char, tar and vo|atiles during pyrolysis is
presented in Fig. 6 [98]. High retention of nitrogen in char and a relation of this
property to coal rank are apparent. The proportion of nitrogen that is evolved
in tar decreases as coal rank increases, which is presumably related to a decrease in volatile matter with coal rank. Another striking feature is a relatively low release of nitrogen as NH3 and HCN: only 6-8%, independent of rank.
Measurements carried out at various temperatures revealed that HCN was
a primary product of coal pyrolysis, with ammonia formed mainly by hydrogenation of hydrogen cyanide. This was concluded from the fact that in most cases
the evolution of HCN started at lower temperatures than evolution of NH3,
and that NH3 was formed simultaneously with H2. These results are consistent
with data by Bose et al. [104] and also by Ghani and Wendt [105] who also found

100

80

F
[]

HCN

[]

NH3

~-N

ch -N

60

40
Z
..&

20

< .......

...........

volatile

coal

matter

rank

........

.........

>

Fig. 6. The d i s t r i b u t i o n of coal-bound n i t r o g e n b e t w e e n t h e char, t a r a n d volatiles d u r i n g


pyrolysis in a n i n e r t a t m o s p h e r e (after B a u m a n n a n d MSller [98]). Coals A F r a n g e in r a n k
from a h i g h - v o l a t i l e b i t u m i n o u s coal to a n a n t h r a c i t e ( c a r b o n c o n t e n t : 81 93% d.a.f., volatile
m a t t e r : 6 40% d.a.f.) Pyrolysis d a t a a v e r a g e d b e t w e e n 973 a n d 1273 K.

22

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

that HCN release preceded NH3 and NO formation during pulverised-coal


combustion. In the case of two coals of the lowest rank studied by Baumann
and M611er, NH3 was found to be formed before H2 release was recorded. This
can be interpreted as a result of direct NH3 formation from amino groups
known to be present in low-rank materials. No appreciable amounts of N2 were
detected in pyrolysis products. A single experiment was performed in which it
was demonstrated that N2 formation could be enhanced by heat treatment at
higher temperatures and longer residence times. The effect of coal type on the
composition of nitrogenous devolatilisation products was also studied by Chen
et al. [106]. They found that, in general, more HCN than NH3 was evolved from
bituminous coals, whereas subbituminous and lignite coals tended to release
larger quantities of NH3 than HCN. This was attributed to variations in the
volatiles content with coal rank.
It seems reasonable to expect that nitrogen functionality in coal plays a role
in the release of NH3 and HCN. As discussed above, amino groups present in
low-rank coals are believed to form NH3 directly. It is unclear if the amount of
amino species is sufficient to account for the observed release of NH3. Thus,
although cleavage of heterocyclic forms of the pyrrolic and pyridinic type is
likely to result in HCN formation, some release of NH3 cannot be ruled out.
More work is needed to elucidate the influence of nitrogen functionality in
coal on speciation of volatile-N. A schematic representation of the fate of
coal-bound nitrogen during pyrolysis is shown in Fig. 7.
The effect of oxygen on pyrolysis of nitrogenous constituents of coal was
studied by carrying out pyrolysis at 873 K with up to 10% oxygen present in the
gas phase [98]. The results render further support to the proposition that NH3
is formed by hydrogenation of HCN. One vol.% of oxygen added to the gas
apparently acts as a scavenger of H2, which results in lower Hi and NH3 levels
than those observed without the presence of oxygen. The concentration of
HCN, on the other hand, is higher but decreases at higher oxygen levels. This

pyrrolic-N
pyridinic-N [J \

CHAR-N
.~ HCN
~\\\\\\\~?

COAL-N

aminogroups

HCN
H2

NH3

Fig. 7. A schematic representation of the fate of coal-bound nitrogen during pyrolysis.

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

23

is explained by HCN oxidation to NO and N2. Char-N and tar-N were found to
be more resistant to oxidation than the other char and tar components.
3.2 Gas-phase reactions
3.2.1 Equilibrium considerations

Let us consider a coal particle undergoing devolatilisation under FBC


conditions. We shall examine a limiting case of flue gas equilibrium composition, assuming that all coal-N is released as NH3 and HCN. Furthermore, let us
assume that the nitrogen and sulphur contents of coal are 1.3% and 1.0%
(d.a.f.) on weight basis, respectively, and that the NH3/HCN ratio in the
volatiles is approximately nine (on a molar basis). The latter value is adopted
on the basis of measurements of NH3 and HCN concentrations at the bottom of
a circulating fiuidised-bed combustor (CFBC) reported by ~ m a n d et al. [107]. In
the same study, the levels of CO at the bottom of the CFBC were found to be
approximately 2%. Amounts of devolatilisation products such as
CH4, CO, CO2, H20 and H2 can be estimated from the data by Suuberg [108]
and Suuberg et al. [101] (a bituminous coal, rapid pyrolysis at 1173 K). This,
combined with typical values for concentrations of other gases, leads to the
following gas composition (vol.%): 0.14%NH3; 0.016%HCN, 12%CO2,
2.0% CO, 0.24% CH4, 0.67% H2, 0.052% SO2, 0.00056% S03, 6.0% H20, 5.6% 0 2
and 73.3% N2.
Multireaction equilibrium composition for the above gas mixture can now be
determined using a classic non-stoichiometric approach based on minimisation
of the Gibbs energy of the system. The other species considered in the reaction
mixture, the concentrations
of which are initially set to zero, are:
NO, N20, NO2, OH, O, H, N, NH, NH2, HNCO, NCO, CN, HNO, S, SH, NS, SO,
COS and H2S.
The results of computations for the species of interest are shown in Fig. 8.
A sharp rise in the concentration of nitric oxide with temperature is striking. It
reflects the increasing contribution from thermal NO at high temperatures but
under FBC conditions, i.e. at about 1123 K, the equilibrium NO concentration
is only 50 ppm. This nearly exactly corresponds to the emission of NO from
a CFBC in which bituminous coal was eombusted, as reported by ~mand et al.
[107]. This agreement may be partly fortuitous, nevertheless the fact remains
that typical NO emissions from FBC's are of the same order of magnitude as the
equilibrium concentration of this species. For example, NO emission levels
observed in 65 and 226 MW FBC units were respectively measured to be 40-260
and 23-40 ppm [109]. The presence of limestone in a CFBC was reported to
increase NO emission above its equilibrium level [110].
The situation is very much different with N20. Figure 8 clearly shows that
the equilibrium concentration of N20 is virtually zero, which thermodynamically favours N20 destruction. Unfortunately, N20 levels observed in practice
are much higher, up to 200 ppm, which implies limitations associated with
reaction kinetics. Unlike with NO, the residence time in an FBC is apparently
insufficient for N20 to approach its equilibrium concentration. Raising temperature could improve N20 destruction kinetics, but the effect on NO would be

24

M.A. Wdjtowiez et al./Fuel Processing Technol. 34 (1993) 1-71

1000

74.5

-ff

NO
800
N2

600

SO3

SO2

74.3
400

0
Z

d
200

dZ
0
5O0

1000
Temperature

10OO

8OO

[K]

Ho-o5

co 2 x

74.1
2OO0

1500

7.5

OH
6.5

600

Ox 5

,-

5.5

H2x2

400 ~

4.5

200
02 [ ~
0
500

'
1000
Temperature

3.5
1500

2000

IKI

Fig. 8. Equilibrium concentrations of selected products of bituminous coal combustion


(initial composition of the gas mixture: 02: 5.6%; N2: 73.3%; CO2: 12%; CO: 2.0%; H20:
6.0%; NH3: 0.14%; HCN: 0.016%; CH4: 0.24%; H2: 0.67%; SO2: 0.052%; SO3: 0.00056%; species
with zero initial concentration: N20, NO, NO2, OH, H, O, N, NH, NH2, HNCO, NCO, CN,
HNO, S, SH, NS, SO, COS and H2S). Equilibrium concentrations of species not shown in the
figures are below 3 ppm in the whole range of temperatures.

adverse. Therefore, c a t a l y t i c e n h a n c e m e n t of N 2 0 d e c o m p o s i t i o n should be


c o n s i d e r e d an a t t r a c t i v e option.
The results of the equilibrium c o m p u t a t i o n s presented a b o v e were found to
be r e l a t i v e l y i n s e n s i t i v e to initial c o n c e n t r a t i o n of H C N and NH3. A s s u m i n g
o n l y 350 ppm of NH3 and 40 ppm of HCN, as m e a s u r e d by A m a n d et al. [107],
m a k e s no q u a l i t a t i v e difference in the equilibrium c o n c e n t r a t i o n s of NO and
N 2 0 . Figure 8 also i n d i c a t e s that under FBC c o n d i t i o n s , the equilibrium
c o n c e n t r a t i o n of NO2 is small and that SO2 is the p r e d o m i n a n t oxide of

25

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

sulphur. This is in agreement with emissions observed in practice. The equilibrium concentrations of H2, CO, O, H and OH rise sharply at temperatures
substantially exceeding FBC combustion temperature.
Similar equilibrium computations were performed for the case of low initial
concentration of oxygen (0.5% instead of 5.6% used previously), which is
supposed to represent conditions prevailing at the bottom of a combustor with
air staging. The results are shown in Fig. 9 and, in general, they bear much
resemblance to what is observed in gasification. In the oxygen-lean environment, formation of H2S rather than SO3 is favoured and the equilibrium
concentrations of H2 and CO are appreciable. High CO levels certainly facilitate

1000

78.2
N O x I0

800

;,

600

so 2

H2S

\f. ox,o, i

4o0

77.8

77.4
z

200

77.0

0~
500

2000

I500

1000

Temperature [KI

1000 l

800i

20

\ H' ~-x l 0

500 i ~ /

"

C~

~001 Jk
2

"~ j
0

o7

500

16

/I

12

o,oo'--4
~
/"
~

1000

= ~o

~2 __.

l-lx i01

Temperature

8
4

1500

2000

tK]

Fig. 9. Equilibrium c o n c e n t r a t i o n s of s e l e c t e d products of b i t u m i n o u s coal c o m b u s t i o n at


reduced o x y g e n level (the same gas mixture as in Fig. 8 w a s used for c a l c u l a t i o n s , except the
a m o u n t of 02 w a s reduced to an initial c o n c e n t r a t i o n of 0.5%). Equilibrium c o n c e n t r a t i o n s
of s p e c i e s not s h o w n in the figures are b e l o w 3 ppm in the w h o l e r a n g e of temperatures.

26

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

reduction of nitrogen oxides. At temperatures characteristic of FBC, the


equilibrium concentrations of all oxides of nitrogen are negligibly low, which
illustrates the effect of excess oxygen on NzO/NOx emissions.
3.2.2 Gas-phase N 2 0 / N O formation and destruction

As mentioned before, N20 chemistry will be discussed along with reactions


responsible for NO formation and destruction. In fact, apart from direct interactions between N20 and NO, there are more than 200 gas-phase reactions
involving species that are implicated in the overall chemical scheme (mostly
radicals, NH3 and HCN). All these reactions have been used in combustion
modelling studies [38, 50 56] and extensive compilations of their kinetic data
are available [38, 50, 52-56]. Detailed discussion of nitrogen chemistry in combustion is beyond the scope of this paper and the reader is referred to a review
on this subject by Miller and Bowman [56]. Only selected reactions will be
discussed here, mostly those that are believed to play a role under fluidised-bed
conditions. A list of these reactions is presented in Table 2.
There are three gas-phase pathways in which NO can be formed during
combustion: thermal-NO, prompt-NO, and fuel-NO. Thermal-NO is formed
according to the Zeldovich mechanism,
N2 + O

N+O2

NO + N
, NO+O

(R17)
(R18)

and, due to relatively low temperature, this route is not considered important
in FBC's. Prompt-NO formation is dominant under fuel-rich conditions, i.e. in
the presence of hydrocarbon radicals, and it is initiated by the reaction,
CH

+Nz

, HCN+H

(9)

and is sustained by the following type of reactions:


CH 2- q-X

CH- + H X

(10)

where X stands for either H or OH. HCN then serves as a nitric oxide
precursor, as discussed below. The relative importance of prompt-NO is rat her
uncertain and, in most cases, it is thought to be small (about 5%) compared
with thermal-NO [111]. In the case of low-NOx burners, however, it is the main
source of NO. Prompt-NO is usually neglected in discussions oriented towards
control technology [112].
Fuel-NO results from oxidation of NH3 and HCN as well as from combustion
of char-N, and is believed to be the main source of NO in fluidised-bed combustion. Ammonia and HCN also happen to be gas-phase precursors of N20, and
the pathways leading to N20 and NO are often competitive.
Houser et al. [44] studied NO/N20 formation from HCN in stirred- and
plug-flow reactors. They concluded that the mechanisms leading to N20/NO
formation were interrelated due to competition for a common gas-phase precursor. It was also found that once N20 is formed at milder temperature conditions, it will not oxidise to NO, even if the reactor temperature should be raised
at a later stage to complete the combustion process. This observation was later
confirmed by Martin and Brown [49] who found that in NzO-doped methane

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

27

and h y d r o g e n flames, N 2 0 c o n v e r s i o n to NO was only a b o u t 5% at tempera t u r e s b e t w e e n 1600 and 2000 K, the r e m a i n i n g 95% being c o n v e r t e d to N2. In
a n o t h e r paper, M a r t i n and B r o w n c o n c l u d e d t h a t N 2 0 f o r m a t i o n was less
t e m p e r a t u r e - d e p e n d e n t t h a n N 2 0 destruction, the l a t t e r also depending on
H radicals c o n c e n t r a t i o n , itself a t e m p e r a t u r e - d e p e n d e n t p a r a m e t e r [54].
E x p e r i m e n t s carried out by K r a m l i c h et al. [48], who injected NH3, H C N and
a c e t o n i t r i l e into the postflame gases, clearly d e m o n s t r a t e d t h a t N 2 0 was
preferentially formed from c y a n o species, w h e r e a s very little NzO was released
d u r i n g NH3 injection (Fig. 10). N 2 0 f o r m a t i o n was observed in a t e m p e r a t u r e
w i n d o w between 1100 and 1500 K. The m a x i m u m for a c e t o n i t r i l e was s o m e w h a t
shifted with respect to H C N due to the delay caused by the droplet e v a p o r a t i o n
time. The following r e a c t i o n sequence leading to NO and N 2 0 f o r m a t i o n from
H C N and NH3 was postulated:
F o r m a t i o n from NH3
NH 3 + OH
, NH2 + H 2 0
N H 2 + OH
NH + NO

(R1)

, NH + H 2 0

(R2)

, N20 + H

(R8)

NH+H

, N+H2

(R4)

N + OH

, NO + H

(R5)

NH + O

, NO + H

(R6)

250

( ' C ~ HCN

~3c~200

>"
15o
el::

C~~

ACETONITRILE

lOO

dcq
z

50

1800

INJECTION TEMPERATURE, K
Fig. 10. Emissions of N20 from a tunnel furnace for NH3, HCN and acetonitrile injection at
various locations (temperatures). Conditions: fuel-lean natural gas combustion (4 = 0.8) with
600 ppm NO; NH3, HCN or acetonitrile added to equivalent of 900 ppm averaged over the
reactor [48]. (Reprinted by permission of Elsevier Science Publishing Co., Inc. from "Mechanisms of nitrous oxide formation in coal flames," by J.C. Kramlich, J.A. Cole, J.M. McCarthy, W.S. Lanier and J.A. McSorley, Combustion and Flame, Vol. 77, pp. 375 384. 1989
by The Combustion Institute).

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

28

Formation from HCN


HCN+O
, NCO+H
HCN + OH

, HNCO + H

HNCO + OH --

(R9)
(R12)

, NCO + H20

(R13)

NCO + NO

N20 + CO

(R14)

NCO+OH

, NO+CO+H

(R16)

Redistribution of HCN- and NH3-based spemes


HNCO + H
~-~ NH2 + CO

(R29)

N20 destruction

N20 + H

, N2 + OH

(R19)

Reaction rate sensitivity analysis shows that reaction (R19) is fully capable
of removing N20 from the flame zone, even if the formation chemistry is
unrealistically augmented. High activation energy for this reaction explains
why hardly any N20 survives high-temperature flame conditions.
The above set of reactions was used to model the experimental results and
good agreement was obtained if reaction (R14) was assumed to branch 40% into
N20 + CO and 60% into N2 + CO2. The following interpretation was given for
N20 release behaviour (Fig. 10): (1) At low temperatures (< 1150 K), reactions
R12 and R29 dominate HCN consumption and the production of the NCO
needed for N20 formation becomes insignificant. (2) Within the N20 formation
window, (R9) and (R14) are predominant. (3) At high temperatures (> 1500 K),
reaction (R16) competes favourably with (R14) for the available pool of NCO.
Also, rapid N20 destruction occurs through reaction (R19).
In summary, at relatively low combustion temperatures, such as in FBC's,
HCN acts as an N20 precursor, whereas at high temperatures, NO is preferentially formed from HCN. This constitutes the essence of the NO-versus-N20
trade-off often observed experimentally.
The above general reaction scheme was later confirmed by other researchers
with occasional modifications resulting from inherent features of the system
studied [38, 43, 53, 56]. For example, Kilpinen and Hupa [53] found that in
addition to reaction (R19), N20 destruction involved OH radicals (R20). They
also argued that inclusion of (R20) into the reaction scheme makes it unnecessary to account for branching of reaction (R14), which they believe does not
take place.
Although only to a minor extent, oxidation of NH3 also results in N20
release [38, 49, 53, 54, 56]. Martin and Brown [49, 54], for instance, studied
concentration profiles of N20, NO and N2 in atmospheric-pressure flat-flame
burner experiments. They found that in an ammonia-doped flame, the main
product was NO (about 70% of fuel-N), with the largest observed concentration
of N20 (in the early-flame region) less than 1% of the NO value. The underlying

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

29

c h e m i s t r y is r e p r e s e n t e d by r e a c t i o n s (R1)-(R3), followed by,


NH~+NO --

, N20+H~

(R7),(R8)

N 2 0 d e s t r u c t i o n is a g a i n by r e a c t i o n (R19).
It is i n t e r e s t i n g to note t h a t no m a t t e r w h e t h e r N 2 0 is formed from H C N or
from NH3, NO is a l w a y s needed to r e a c t with either NCO or NH~. It is n o t
s u r p r i s i n g t h e n t h a t a s t r o n g e n h a n c e m e n t of N 2 0 f o r m a t i o n was observed by
H u l g a a r d [38] upon addition of NO to a gas m i x t u r e d u r i n g NH3 and H C N
o x i d a t i o n (Fig. 11). E x p e r i m e n t s were carried out in a plug-flow q u a r t z reactor.
F i g u r e 11 shows t h a t only small a m o u n t s of N 2 0 are formed in the presence of
NH 3, w h e r e a s c o n v e r s i o n of H C N to N 2 0 is m u c h larger. A shift in N 2 0
c o n c e n t r a t i o n profiles caused by the presence of CO can be explained as the
result of an increased c o n c e n t r a t i o n of radicals, w h i c h in t u r n leads to enh a n c e d N 2 0 f o r m a t i o n and destruction.
Kinetics of gas-phase N 2 0 d e c o m p o s i t i o n a c c o r d i n g to r e a c t i o n s (R19) (R21)
h a v e been r e c e n t l y studied in t u b u l a r q u a r t z r e a c t o r s by H u l g a a r d [38] and
K h a n et al. [113]. They found a p p a r e n t a c t i v a t i o n e n e r g y for this process to be
241 ( H u l g a a r d ) and 247 kJ/mol ( K h a n et al.). Only up to a b o u t 2% of N 2 0 was
c o n v e r t e d to NO and this effect was s t r o n g e s t at high t e m p e r a t u r e s ( ~ 1400 K)
[381.

60

~x~x

x\

50

e-i

40

/f

z
m

"?\..

9O0

1100
Temperature

1300

i500

[K]

Fig. 11. N20 formation during gas-phase oxidation of HCN and NH3 (Hulgaard, 1991 [38]).
Experiments carried out in a quartz plug-flow reactor at a constant flow rate. Residence time
ca 55 ms at 1200 K. Legend: (C;) NH3; (O) HCN; (A) NH3/NO; (A)HCN/NO; (+) HCN/CO;
and (x) HCN/NO/CO. (Inlet concentrations for HCN runs: HCN: 310 ppm, 02: 2.3%; H20:
2.7%; NO: 440 ppm; CO: 1750 ppm; N2: balance. Inlet concentrations for NH3 runs: NH3:
770 ppm; 02: 2.6%; H20: 1.0%; NO: 600 ppm; N2: balance).

30

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

Simulation of homogeneous formation and destruction of nitrogen oxides at


conditions typical for FBC was carried out by Grimsberg and Karlsson [52].
The model used in that study comprised 87 reactions and 26 species. A strong
temperature dependence for N20 reduction was found (Fig. 12), which is in
agreement with the data discussed above. A strong dependence of N20 destruction on residence time should also be noted. The effect of excess oxygen was
found to be weak in the range between 1 and 5% but, again, increasing
residence time would have a beneficial effect on reducing N20 emission.
Modelling of gas-phase reactions in an FBC was also performed by Kilpinen
and Hupa [53]. They used a detailed kinetic scheme comprising 49 species and
253 reactions, out of which 17 directly involved N20. This was combined with
an ideal plug-flow reactor model. Two types of fuel were used in calculations:
methane and a typical fuel-rich coal combustion flue gas but the model turned
out to be relatively insensitive to fuel type. The results show that HCN
produces significantly higher concentrations of N20 as compared with NH3
(N20 yield was found to be nearly linear with the initial HCN concentration).
At lower temperatures, substantially higher amounts of N20 were produced.
The distribution of nitrogenous products of combustion is presented in Fig. 13.
Most of the HCN is oxidised to NO at high temperatures, whereas N20 is the
predominant oxide of nitrogen at low temperatures. Conversion to N2 is also
favoured at low temperatures. Unlike in the work by Grimsberg and Karlsson,
the effect of excess air ratio was found to be pronounced at 1100 K. Hardly any
N20 is formed at stoichiometric ratios below unity but for stoichiometric
ratios greater than unity N20 emission level rises sharply and then levels off
(Fig. 14). Similar calculations were carried out for 1200 K and the results show
very low N20 levels irrespective of stoichiometric ratio. This may be due to

2IZ
u

0
1000

1100

1200

1~100

Temperature [K]

Fig. 12. Gas-phase NzO destruction as a function of temperature and residence time (Grimsberg and Karlsson, 1990 [52]). Oxygen concentrations: 5%.

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

l!

31

100

~0
1

[~

N2

I~

NO2
NO

[]

N20

HCN

[]

HNCO

4O
.-2

120

01

L'

' ,~',~

1000

1100

1200

Temperature

1300

[K]

Fig. 13. Model p r e d i c t i o n s for n i t r o g e n d i s t r i b u t i o n a m o n g p r o d u c t s of H C N o x i d a t i o n [53].


I n i t i a l HCN c o n c e n t r a t i o n in coal c o m b u s t i o n flue gas: 500 ppm; s t a g e d c o m b u s t i o n
primary s t o i c h i o m e t r i c r a t i o = 0.7; overall s t o i c h i o m e t r i c r a t i o = 1.2. (Reprinted by p e r m i s s i o n
of Elsevier Science P u b l i s h i n g Co., Inc. from " H o m o g e n e o u s N 2 0 c h e m i s t r y at fluidized bed
c o m b u s t i o n conditions: A k i n e t i c modeling study," by P. K i l p i n e n a n d M. Hupa, C o m b u s t i o n
a n d Flame, Vol. 85, pp. 94 104. (C 1991 by The C o m b u s t i o n Institute).

100

c~ ppmv
g
a0
1100 K
z0

60

CI;

L,O

20

1200 K
0,9

1,0

1,1

1,2

1,3

1,/,

1,5

1,6

STOICHIOMETRY
Fig. 14. C a l c u l a t e d N 2 0 emission versus air fuel s t o i c h i o m e t r i c r a t i o [53]. C o m p u t a t i o n s
were b a s e d on plug-flow c o n d i t i o n s at 1100 and 1200 K, at a residence time of 3.0 s. M o d e l i n g
was performed a s s u m i n g m e t h a n e doped w i t h 0.6 tool% HCN as a fuel. (Reprinted by
p e r m i s s i o n of Elsevier Science P u b l i s h i n g Co., Inc. from " H o m o g e n e o u s N 2 0 c h e m i s t r y at
fluidized bed c o m b u s t i o n conditions: a k i n e t i c modeling study," by P. K i l p i n e n a n d M. Hupa,
C o m b u s t i o n a n d Flame, Vol. 85, pp. 94 104. ((~ 1991 by The C o m b u s t i o n I n s t i t u t e . )

32

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

rapid N20 d e s t r u c t i o n t a k i n g place at 1200 K. Finally, the effect of air staging


was examined. In these c o m p u t a t i o n s , the p r i m a r y stoichiometric ratio was
varied between 0.6 and 1.2, whereas the overall s t o i c h i o m e t r i c ratio was kept
c o n s t a n t at 1.2. The fuel was modelled as m e t h a n e doped with 0.6% HCN.
Results show t h a t final NEO c o n c e n t r a t i o n is v i r t u a l l y i n d e p e n d e n t of the
p r i m a r y s t o i c h i o m e t r i c r a t i o in the r a n g e 0.6-0.9. At h i g h e r values, however,
N20 emission declines to a m i n i m u m at a b o u t 1.0 and r e t u r n s to its original
level at a p r i m a r y s t o i c h i o m e t r i c ratio of 1.2. This is in a g r e e m e n t with the
e x p e r i m e n t a l d a t a by K r a m l i c h et al. [114].
K i l p i n e n and H u p a also identified and quantified the most i m p o r t a n t reaction p a t h w a y s for H C N oxidation u n d e r FBC conditions. F o r m a t i o n of NCO
was found to be governed m a i n l y by r e a c t i o n (R9) but also to a lesser e x t e n t by
r e a c t i o n s (R10) and (Rll):
HCN + O --

~ NCO + H

H C N + OH

, CN

CN+O 2

+ H20

NCO+O

(R9)
(R10)
(Rll)

D e s t r u c t i o n of NCO depends on local c o m b u s t i o n conditions and it m a y follow


the following four routes.
(a) NCO + H
(b) NCO +
NCO +

, N H + CO

H20
H 2 --

(c) NCO + O

HNCO + OH

HNCO + H

> NO + CO

NCO+OH-(d) NCO + NO

(R25)
(R26)
(R27)
(R15)

, NO+CO+H

(R16)

, N 2 0 + CO

(R14)

At high t e m p e r a t u r e s ( ~ 1200 K), routes (a) and (c) prevail, r o u t e (a) being
followed by NO f o r m a t i o n t h r o u g h r e a c t i o n s (R4), (R5) and (R18):
NH + H

~ N + H2

(R4)

N + OH

-+ NO + H

(R5)

N+O2

, NO+O

(R18)

This of course results in fast NO formation. At lower t e m p e r a t u r e s


( ~ 1000 1200 K), routes (a) and (c) are slow e n o u g h so t h a t N 2 0 f o r m a t i o n may
o c c u r via r o u t e (d). U n d e r fuel-rich conditions, h a r d l y any N 2 0 is formed since
H C N does not decompose to form NCO due to low c o n c e n t r a t i o n of radicals.
The small a m o u n t of H C N t h a t does r e a c t to NCO forms mostly NH3 a c c o r d i n g

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

33

to the following pathways:


H2 H20

(b)

NCO

~ HNCO

(R26), (R27)

, NH2

(R29)

HNCO
H2 H20

NH2

, NH3

(11)

At low temperatures ( < 1000 K), the rate of HCN decomposition drops, which
leads to slow N20 formation. Finally, N20 destruction is governed by reactions
(R19) and (R20), both important in the 1000 1200 K temperature range. Increased N20 levels at high oxygen excess are interpreted as a result of inhibition of R19 due to a shortage of H radicals.
De Soete [69] suggested that NOx may interact with SOx in an FBC in a way
similar to that responsible for the N20 sampling artifact. So far very little has
been done to explore this possibility. Some preliminary data have been collected [115, 116] which show that introducing SO2 into a laboratory-scale
fluidised bed (fired with methylamine-doped propane) causes redistribution of
nitrogenous combustion products. More N20 is formed and less NO when SO2
is present in the system. Similar results were obtained using pyridine instead of
methylamine [117].
The effect of SO2 on NO is not surprising in view of the results of premixed
flame experiments carried out in 1970's and early 1980's. Wendt and Ekmann
[118] used a methane-air flat flame doped with either SO2 or H2S to show that
sulphur inhibits formation of NO. They attributed this result to catalysis of
oxygen atom recombination reactions by SO2 and concluded that fuel desulphurisation might lead to increased NO~ emissions. It was also shown that H2S
must be converted to SO2 before NO inhibition is important. Such a conversion
occurs readily under fuel-lean conditions.
Tseregounis and Smith [119] studied fuel-rich, premixed hydrogen and acetylene flames and concluded that the effect of sulphur compounds on NO occurred via a direct or indirect interaction with NHi radicals. Corley and Wendt
[120] used premixed CI-I4/He/O2 flames and found that SO2 injection resulted in
reduced NO, N2 and NH3 levels, whereas the concentration of HCN increased
in the postflame region as compared with the reference case. This effect was
explained by the interaction of SO2 with the cyanide and amine subsystem.
In a different study, in which CO/Ar/O2 flames were used, Wendt et al. found
that the effect of sulphur was to decrease postflame NO levels and to increase
N2 concentration [121]. Modelling, which involved over 60 reactions, showed
that the most plausible effect of SO2 was to increase the steady-state concentration of nitrogen atoms, and consequently to enhance N2 formation. This
occurred via direct interactions between N, NO, S and SO.
Finally, it should be pointed out that H, OH and/or O radicals are necessary
to form NO from NH3- and HCN-related species (reactions R5, R6, R15 and

34

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

R16). SO2 may catalyse radical recombination, leading to a decrease in NO


levels upon SO2 injection, as discussed below.
Webster and Walsh studied inhibition of explosions in hydrogen and oxygen
mixtures caused by SO2 and found evidence for the following reaction [122]:
H + SO2 + M

, HSO 2 +

(12)

where M stands for a gas molecule. To account for the enhanced rate of H and
OH recombination in SO2-containing flames, Fenimore and Jones [123] proposed a mechanism involving the above reaction followed by the rapid bimolecular steps:
H + HSO2
OH + HSO 2

, H2 + SO2
, H20 + 802

(13)
(14)

In the above scheme, HSO2 clearly acts as a scavenger of H and OH radicals.


This mechanism was later confirmed by Kallend [124], who studied chemiluminescence in hydrogen flames, as well as by Halstead and Jenkins [125] who
measured hydrogen atom concentrations in h y d r o g e n + o x y g e n + n i t r o g e n
flames.
Only in one of the above-mentioned studies was the presence of N20 reported
in the postflame zone but this effect was not quantified [120]. It is likely that an
increase in NzO levels following SO2 injection also results from a reduced pool
of H and OH radicals. These radicals are instrumental in N20 destruction via
reactions (R19) and (R20). The effect of SO2 on N20/NO emission in an FBC is
discussed further in Section 4.1.
3.3 Heterogeneous reactions
3.3.1 Heterogeneous N 2 0 / N O formation and destruction on char surface
Potential sources of N 2 0 originating from char-N include the following:
(1) The oxidation of char-N to N20 (reactions R30, R33 and R34).
(2) The reaction of NO with char nitrogen (R39), or the adsorption of an NO
molecule onto char surface where it subsequently reacts with anot her NO
molecule (e.g. reactions R38a and R39).
(3) The gasification of char-N to form HCN or NH3, which is followed by
gas-phase HCN/NH3 oxidation to N20/NO.
Kramlich et al. [48] discounted the first possibility as mechanically unlikely
(two neighbouring N-containing sites are required) and modelled formation of
N20 according to the other two pathways. The results indicated that route (2)
would result in unrealistically low N20 concentrations, even if all NO-reduction on char surface led to N20 as a product. Modelling of char-N release as
HCN, however, resulted in an N20 emission of approximately 100 ppm. Thus it
was concluded that route (3) was most likely to describe char-N conversion to
N20.
Heterogeneous reactions of NO and N20 on highly devolatilised char surfaces were studied by de Soete [45, 65], whose model included only char-N
oxidation to N20/NO (i.e. route (1)). More specifically, the kinetic model

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 7l

35

c o m p r i s e d r e a c t i o n s (R30), (R31), (R33), (R38) and R(42), in a d d i t i o n to c a r b o n


o x i d a t i o n to CO and CO2. U s i n g b o t h s t e a d y - s t a t e and t r a n s i e n t e x p e r i m e n t s in
a l a b o r a t o r y p a c k e d - b e d r e a c t o r , k i n e t i c r a t e c o n s t a n t s w e r e d e t e r m i n e d for
s e v e r a l chars. De Soete verified e x p e r i m e n t a l l y t h a t u n d e r the e x p e r i m e n t a l
c o n d i t i o n s used in his work, the c o n t r i b u t i o n to N 2 0 / N O f o r m a t i o n c o m i n g
f r o m r o u t e (3) w a s small (3 8%). Results of his s t u d y show t h a t d u r i n g c h a r
c o m b u s t i o n , N 2 0 and NO are m a i n l y r e l e a s e d as d e s o r p t i o n p r o d u c t s f r o m
oxidised c h a r - N atoms. The f r a c t i o n s of c h a r - N t r a n s f o r m e d into N 2 0 a n d NO
w e r e f o u n d to be p r o p o r t i o n a l to c h a r burn-off, w h i c h m e a n s t h a t c h a r - b o u n d
c a r b o n a n d n i t r o g e n a t o m s are oxidised in p r o p o r t i o n s n e a r l y i d e n t i c a l to t h e i r
r e s p e c t i v e c o n c e n t r a t i o n s . It was also d e m o n s t r a t e d t h a t NzO is m o r e r e a d i l y
r e d u c e d on c h a r s u r f a c e t h a n NO (Fig. 15). The r a t i o of o v e r a l l N 2 0 / N O
r e d u c t i o n r a t e c o n s t a n t s at 1123 K ( m e a s u r e d in the a b s e n c e of r e d u c i n g gases
s u c h as CO) was found to be 7.7, 7.0, 15 and 13 for g r a p h i t e , C e d a r G r o v e char,
P r o s p e r c h a r and E s c h w e i l e r char, r e s p e c t i v e l y . T r a n s i e n t e x p e r i m e n t s rev e a l e d t h a t o v e r a l l N 2 0 / N O f o r m a t i o n was c o n t r o l l e d by o x y g e n a d s o r p t i o n .
T y p i c a l l y , 1 - 6 % of c h a r - N was found to be c o n v e r t e d to N20. Wdjtowicz et al.
[126] r e p o r t e d a s i m i l a r r e s u l t in l a b o r a t o r y - s c a l e FBC e x p e r i m e n t s (up to 10%
c h a r - N c o n v e r t e d to NzO).
In a different s t u d y [127], De Soete f o u n d t h a t f o r m a t i o n of N 2 0 d u r i n g NO
r e d u c t i o n on c h a r s u r f a c e ( r e a c t i o n s R38 and R39) was negligible at tempera t u r e s below 1300 K, u s u a l l y below d e t e c t i o n limits. Similar e x p e r i m e n t s were
c a r r i e d out in a p a c k e d - b e d r e a c t o r by B a u m a n n [128] and in a l a b o r a t o r y - s c a l e

-12

-14
"7,
Z

-16
_=

-18

0'.8

i'.o
liT x 1000

-7:2

1.4

[K]

Fig. 15. Reaction rate constants for heterogeneous reduction of N20 and NO on Cedar Grove
char (De Soete [45, 65]). Particle size: 35 -50 pro; inlet gas composition: () 80 ppm N20/Ar;
(~)) 80 ppm N20/0.46% CO/Ar and (Q) 57 ppm NO/0.24% CO/Ar.

36

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

FBC by Gulyurtlu et al. [129, 130] and Mochizuki et al. [131]. Inlet concentrations of NO were about 1000 ppm in all these studies and no significant
amounts of N20 were detected at the reactor outlet (N20 was usually below the
detection limits of 1 ppm [128] and 10 ppm [131], and less than 1% of the reduced
NO [129, 130]). In the work by Baumann a gas mixture containing up to 1% CO
was also used and no N20 was detected in the reaction products. Implications
of this important result for fluidised-bed combustion are discussed in Section 4.
It was also reported that N20 reduction on graphite (reaction R44) gives
rise to formation of only small amounts of NO (less than 1% of the reduced
N20 reacts towards NO; temperature < 1200 K, 77 ppm NzO/Ar and 77 ppm
N20/0.43% CO/0.97% CO2/Ar used in a packed-bed reactor) [127].
Kinetics of NO reduction on char surface were reported by Johnson [67].
A simple power law was used to represent kinetic data for the following
reaction:
NO + C

,N2 q- CO

(15)

The reaction was found to be 0.5-0.6 order in NO with an activation energy of


about 90 kJ/mol.
3.3.2 Heterogeneous N z O / N O formation and destruction on catalytic surfaces
Formation of N20 during solid-catalysed ammonia oxidation was studied by
Iisa et al. [132] in a quartz packed-bed reactor. Nitrous oxide was formed
from ammonia on calcined limestone in the temperature range 1023-1223 K,
provided a sufficient amount of oxygen was present (at least 0.5% 02). The
outlet N20 concentration increased as temperature increased, went through
a maximum, and then decreased. This behaviour indicates that at higher
temperatures, N20 decomposition successfully competes with N20 formation
(see below). The reaction products were N2, N20 and NO; N2 was found to be
the main product at high temperatures. No N20 was detected in the products
when quartz sand and sulphated limestone were used as bed materials.
In the same study, solid-catalysed N20 decomposition was also examined
over beds of quartz sand, calcined limestone, sulphated limestone and peat ash.
All the materials enhanced N20 destruction as compared with NzO decomposition in an empty reactor, calcined limestone exhibiting the strongest effect.
The overall activation energies for this process were found to be 59, 108 and
190 kJ/mol for calcined lime, quartz sand and an empty reactor, respectively.
The N20 conversion was independent of inlet concentration, which indicates
a first order reaction. Adding 10% O2 and, in a separate experiment, 1% CO to
the inlet gas had practically no effect on N20 conversion. Nitrogen gas was the
only reaction product detected, except for the experiment in which CO was
present in the inlet gas. In that case, CO2 was additionally formed. Saturation
of the inlet gas with water vapour slightly enhanced N20 decomposition above
1073 K and somewhat inhibited it at lower temperatures. The authors concluded that under FBC conditions, the effect of N20 decomposition is likely to
be stronger than N20 formation from Ca-catalysed oxidation of ammonia.
Decomposition of NzO on solid surfaces was also studied by Khan et al. [113],
Miettinen et al. [133], W6jtowicz et al. [134], Naruse et al. [135], Kimura [136]

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

37

and Klein et al. [137]. The results are, generally speaking, similar to those
reported by Iisa et al. It is worth mentioning that the catalytic activity of
Fe203 towards N20 decomposition was found to be very strong, comparable
with that of CaO, and also that MgO is quite effective [133, 137], although not
as much as CaO and Fe203 [133].
Reaction pathways underlying N20 decomposition on solids were reviewed
by Golodets [138] and the following two mechanisms were proposed for the
reactions on metallic catalysts (e.g. Pt, Pd, Ag, Au, Ge) [138, 139]:
N 2 0 + (*)

, N2 + (O*)

(16a)

' N2 + 0 2 + (*)

N20 + (0")

(16b)

() fast

02 +(*)

, (02*)

2N20 + 2(0*)

, 2(0*)

, 2N2 + 202 + 2(*)

(17a)
(17b)

Both mechanisms give rise to the same overall stoichiometry:


2N20

, 2N2 + 02

(18)

and both are believed to be necessary for an adequate description of the


reaction kinetics. It should be noted that routes (16) and (17) differ only in the
first step, i.e. in the way surface sites are oxidised. The source of adsorbed
oxygen in (16) is N20, whereas (O*) results from 02 adsorption in route (17). In
a limiting case of oxygen-lean environment, kinetic scheme (17) is eliminated
and the reaction proceeds entirely via route (16). Under oxygen-rich conditions, however, 02 turns out to be a more effective oxidant than N:O and route
(17) prevails. Analysis of kinetic schemes (16) and (17) leads to the following
rate expressions:
2
klAklBPN2o
F(16) = (klA "4-klB)PN:o + 2k2APo2

k l Bk 2APN2oPo2

r(l 7) - (kl A+ k 1B)PN~O+ 2k 2APo2

(19)

(20)

where klA, klR and k2A are rate constants of reactions (16a), (16b) and (17a),
respectively; PN2Oand Po2 stand for partial pressures of N20 and 02. Equations
(19) and (20) have been demonstrated to provide good description for experimental data [139].
The mechanism of N20 decomposition on metal oxides was proposed by
Stone [140]. According to Golodets [138], it can be reduced to two schemes:
scheme (16) and (21) described below:
N20 + (*) ,~ (N20*)

(21a)

(N20*)

(21b)

2(0*) ~

, (O*)+ N 2

02 + 2(*)

(21c)

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

38

W i t h s o m e s i m p l i f i c a t i o n s , t h e f o l l o w i n g r a t e e q u a t i o n c a n be d e r i v e d for
(21):

r~ =

k'PN2
l+~o2Po

(22)
2

w h e r e k' is a r a t e c o n s t a n t a n d bo2 is t h e a d s o r p t i o n coefficient for o x y g e n .


B o t h r a t e e x p r e s s i o n s (19) a n d (22) a r e c o n s i s t e n t w i t h t h e e x p e r i m e n t a l
o b s e r v a t i o n s t h a t : (a) t h e r a t e is u s u a l l y p r o p o r t i o n a l to PN2O [141 147], a n d
(b) t h e p r o c e s s is o f t e n r e t a r d e d b y o x y g e n [142 144, 146].
T h e c a t a l y t i c a c t i v i t y of s e l e c t e d m e t a l o x i d e s is p r e s e n t e d i n T a b l e 4. T h e
o x i d e s a r e a r r a n g e d i n o r d e r of d e c r e a s i n g a c t i v i t y , d e f i n e d a r b i t r a r i l y as

TABLE 4
Catalytic activity of metal oxides towards N20 decomposition (after Golodets [138]). k is
a reaction rate constant, defined by the equation r =- kPN2o/x/Po2; k = A e x p ( - E / R T ) To is the
lower temperature at which reaction could be observed
Catalyst

Logk
A
E
Temperature
(at723K)
( P a l / 2 s - l m 2) (kJmol 1) range(K)
[Pa ~/2s -1 m 2] Ref. [142]
Ref. [142] Ref. [142]
Ref. [142]

Rh203
Ir02
Co3Q
CaO
CuO
SrO
HfO2
Fe203
NiO
Th02
Mn02
CdO
SnO2
CeO2
ZrO2
MgO
Cr203
ZnO
Nd203
Ga203
BeO
A1203
TiO~
Sb204
WO3
U30s
SiO 2
Ge02

0.22
-0.57

3.0E + 10
3.9E +08

142
128

553 653
523-623

- 1.97
-2.08
-2.47
-2.56
-2.70
3.00
-3.27
3.33

1.7E +08
1.3E +05
2.8E +04
2.9E +06
8.1E +03
7.2E +08
5.5E +05
4.8E + 05

142
100
96.2
126
92.0
165
126
126

443--653
613 683
693 813
753 813
553 733
673 773
693 773
683 753

-3.72
-3.88

7.8E +05
8.5E +03

134
109

703 823
673-773

-4.44
-4.70
-4.99

1.2E +06
2.0E +07
2.1E + 07

146
167
172

723 823
723--843
803-883

-5.55
-5.59
-5.97
-6.18

4.2E +06
3.3E +05
8.5E +06
3.4E + 05

170
155
180
163

813 913
823-933
803-933
813-953

E
To
(kJmol t) (K)
Ref. [141] Ref. [141]

66.9
79.5
75.3

553
913
673

75.3
75.3
71.1

773
567
693

113
117
79.5
79.5
109
117
105
134

753
768
758
758
878
823
823
848

79.5
92.0

948
743

109
87.9
96.2
92.0
100

913
933
948
968
1008

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

39

reaction rate constant k at 723 K. The values of k were calculated using the
following equation [142]:

=k~

(23)

~ y
~/ 02

Another index of catalyst activity was used by Saito et al. [141] who determined the lowest temperature To at which appreciable N20 decomposition
could be observed. The values of To as well as reported activation energies are
also included in Table 4.
It can be seen from Table 4 that the oxides of group VIII transition metals
(Rh, Ir, Co, Fe and Ni) as well as CaO and SrO exhibit the highest activity. The
oxides group III-VII metals (Mn, Ce, Th, Sn, Cr) are moderately active, whereas the oxides of main subgroup metals (A1, Ga, Be, Sb, Si, Ge) exhibit lower
activities.
Calcium oxide is a catalyst of particular interest because limestone is often
utilised to control SOx emission in an FBC. A study of catalytic reactions
involved in the conversion of nitrogenous species on the surface of calcined
limestone was performed by De Soete and Nastoll [147]. They used a laboratory
fluidised-bed r eact or to determine the formation and destruction rates of NzO
and NO, as well as rates of NH3 and HCN decomposition on calcined limestone.
To facilitate kinetic analysis, the number of gaseous reactants was progressively increased: NO, NO/O2, N20, N20/O2, NH3, NH3/NO, NH3/N20, NH3/O2,
NH3/NO/O2, HCN, HCN/NO, HCN/N20, HCN/O2 and HCN/O2/NO. The following conclusions were drawn:
(1) Most CaO-catalysed reactions occur within the temperature range
700-1200 K, which makes them relevant to fluidised-bed combustion.
(2) The presence of 02, NO, NHa or HCN in the gas phase has no effect on
CaO-catalysed N20 decomposition.
(3) Apart from transient formation of small amounts of N20, no significant
amounts of nitrous oxide are formed from NO or NO/O2.
(4) Neither N20 nor NO is formed from HCN alone. In the presence of oxygen,
NO is the main product of HCN oxidation, whereas N20 is formed as
a secondary product from HCN and NO. NzO is also formed from HCN and
NO, even if oxygen is absent from the system.
(5) Qualitatively similar behaviour of CaO and V205 was noted in catalytic
conversion of nitrogen-containing species.
The following set of reactions was proposed to describe the nitrogen chemistry
on the surface of calcined limestone [147]:
N20 destruction
N20 + (Ca)
, N2 + (CaO)

(24)

N20 + (CaO)

(25)

, N 2 + 02 + (Ca)

40

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) I 71

NO destruction
NO + 2(Ca)
2(CAN)
NO + (CAN)

, (CaO) + (CAN)

, N2 + 2(Ca)
, N20 + (Ca)

(26)
(27)
(28)

(also N20 formation)


NO formation
(CaO) + (CAN)
02 + (CAN)

, NO + 2(Ca)
, NO + (CaO)

(29)
(30)

Surface oxygen stripping


2NH3 + 3(CaO)
, 3H20 +2(CAN)+ (Ca)

(31)

Surface oxidation
O2 +2(Ca)
, 2(CaO)

(32)

Elementary steps for reactions (30) and (31) were suggested and kinetic data
for the C a O / N 2 0 / N O / N H 3 / H C N / 0 2
reacting system are also available [147]. It
should be noted that reactions (24) and (25), which represent N20 destruction
on the surface of limestone, are identical to the general reaction scheme (16)
proposed for N20 decomposition on metal oxides. Furthermore, according to
De Soete, reactions (24) and (25) are equally fast at the steady state, which
explains the experimentally observed overall stoichiometry: -AN20/AO2 = 0.5.
The dual reaction path is also consistent with the fact that N20 decomposition
on limestone is affected neither by the presence of oxygen donors (02, NO) nor
by the presence of oxygen stripping species (NH3, HCN). It is not clear at the
moment whether and in what way CO2 affects N20/NO chemistry on the
surface of limestone.
The NO reduction pathway (26)-(27) is autoinhibiting due to a buildup of
(CaO) on the surface. Thus, these reactions may proceed only in the presence of
oxygen-stripping species such as NH3, HCN, and - - to a minor extent - - also
N20 (via reaction 25). Decomposition of NH3, on the other hand, requires an
oxygen donor to keep the inventory of (CaO) high enough for reaction (31) to
proceed. 02, NO and N20 may act as oxygen sources for the surface through
reactions (30), (26) and (24), respectively. The gaseous products of NH3 oxidation are NO (reaction 31 followed by 29 and 30), H20 (reaction 31) and N20
(reactions 31 and 28).
Results of the first studies of catalytic reactions involving N20, NO, 02 and
hydrocarbons have recently been made available. Calcium-oxide-catalysed
N20 reduction was found to be appreciably enhanced by the presence of
hydrocarbons [148]. Formation of N20 during heterogeneous NO reduction in
the presence of hydrocarbons was carried out using a number of metal oxides,
zeolite catalysts and CaO [149]. The results show that within the optimum
temperature range for NO reduction, the conversion of NO to N20 varies
between 0 and 6%, depending on the catalyst, C/NO ratio and oxygen concentration. For more information, the reader is referred to the review by De Soete
[150] and to the references quoted therein.

41

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

A l t h o u g h N 2 0 f o r m a t i o n from NO was found to be of only m i n o r i m p o r t a n c e


on limestone, nitric oxide r e d u c t i o n on o t h e r c a t a l y s t s m a y be a sizeable source
of N20. The problem is well-known in c a t a l y t i c c o n v e r t o r s ( P t - R h ) used to
c o n t r o l car e x h a u s t gases [46] but very little w o r k has been done to elucidate
the role played by such r e a c t i o n s in a coal c o m b u s t i o n e n v i r o n m e n t . De Soete
r e p o r t e d the results of experimental w o r k on N 2 0 f o r m a t i o n d u r i n g decomposition of NO over a reduced C a S 0 4 surface ("CaSOR") [69]. Packed-bed experiments involved d e s t r u c t i o n of N 2 0 / N O on g r a p h i t e and g r a p h i t e mixed with
CaS04. It was found t h a t N 2 0 f o r m a t i o n a c c o m p a n i e d NO r e d u c t i o n w h e n
C a S 0 4 was present in the bed (Fig. 16) and the following r e a c t i o n was proposed
to a c c o u n t for this effect:
2NO + " C a S O R "

, N 2 0 + CaSO 4

(R49)

"CaSOR" denotes a s u l p h a t e r e d u c t i o n product, e.g. CaSO3, CaSO 2 or CaS. The


r e a c t i o n was reported to follow the first-order kinetics in NO, with an activation e n e r g y of 16.6 kJ/mol. Kinetic d a t a show t h a t r e a c t i o n (R49) is f a v o u r e d by

NOouf/NOin
1000

1.00

--8

0.75

750

A
eq

0.50

500

\B

N20 100
0.25

250

lO00

1100

1200
Temperature

1300
[K]

Fig. 16. Formation of N20 as a result of NO decomposition on CaSOR (after De Soete, 1989
[69]). Gas composition: 1240 ppm NO/5 ppm N20/Ar; bed material: (A) 0.1 g graphite +0.5 g
CaSO4, and (B) 0.1 g graphite.

42

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

low temperature, low oxygen partial pressure and high sulphate concentration. Under some conditions, this reaction route may successfully compete
with N20 formation from char N. It was suggested that other sulphates
(e.g. those of alkali metals) may also exhibit a similar, or even stronger activity
[69].
Kinetics of solid-catalysed NO formation and destruction according to reactions (R36), (R45) and (R47) were studied in a differential packed-bed reactor by
Johnsson [67]. The solids used in that work were char, ash and sand and the
study also included ammonia decomposition to N2 over the same solid materials. A simple power-law expression was used for the kinetic description.
Reduction of NO by CO (reaction R45) was found to be strongly catalysed by
char, especially by brown-coal char, the ash of which contained a high content
of CaO and Fe203. The reaction orders in NO and CO varied with temperature,
implying a complex chemical scheme. The catalytic reduction of NO by NH3
was possible only when oxygen was absent. When oxygen was present, ammonia was oxidised to form NO and N2. In general, large differences in the
catalytic activity of different materials (e.g. various kinds of char) were reported, the conclusion being that it is necessary to base modelling of heterogeneous chemistry on kinetic data collected for specific materials. Unfortunately, no measurement of N20 concentration was performed in this study and
thus the importance of reaction (R49) cannot be assessed.
Formation and destruction of N20 during limestone sulphation in alternating reducing/oxidising conditions was studied by Hulgaard [38]. Results of
this work are particularly relevant to staged combustion in CFBC's, where
limestone particles are constantly exposed to periodic changes in reducing and
oxidising conditions. Under such a r6gime, the chemical and physical characteristics of limestone vary periodically (CaO
,CaSO4
,(CaS?)
,CaO,
etc.; pore narrowing and/or closure), which results in changes in the efficiency
of sulphur capture. An experimental study of limestone N20 chemistry under
such conditions was carried out in a packed-bed reactor with a limestone bed
and synthetic flue gas flowing through it. Reducing/oxidising conditions were
implemented by periodic switching from 02 to CO and back to 02 at the reactor
inlet. To monitor the response of the system to such changes, the composition
of the gas mixture was measured at the outlet.
The results show that:
(1) Small amounts of N20 can be formed as the result of NO reduction over
sulphated limestone (i.e. under reducing conditions when CaSOR or CaS are
thought to be predominant on the surface).
(2) In a reducing environment (CO), the presence of SO2 clearly favours
NO reduction over limestone. A rapid decline in the reduction efficiency
is observed when the flow of SO2 is switched off. Reduction of NO
resumes at the original level upon reintroduction of SO2. It is effectively
reduced over limestone, even when there is no SO2 present in the reaction
mixture.
(3) Decomposition of N20 takes place on both CaO and CaS. Reducing conditions (CO, CaS) strongly enhance this process and little decomposition
occurs under oxidising conditions (02 instead of CO).

M.A. WOjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

43

(4) Sulphur dioxide inhibits N20 decomposition unless CO is present in the


system. In the latter case, the presence of SO2 has no adverse effect on N20
reduction.

4. N20/NO FORMATION AND DESTRUCTION IN FLUIDISED-BED COMBUSTORS


4.1 Effect of fuel type and operating conditions
4.1.1 Temperature

The effect of temperature on N20 and NO emissions from an FBC is perhaps


the most obvious of all and one about which there is no controversy in the literature. As discussed in Section 3, there exist powerful gasphase mechanisms for N20 removal, either by radicals or through collisions
with gas molecules. In addition, heterogeneous destruction of N20 can also be
sizeable, although its contribution certainly depends upon the concentration
of particles. In principle, if given enough residence time in a combustor, N20
would eventually reach its equilibrium concentration, which happens to be
nearly zero. Since the residence time is finite, temperature is clearly the most
important factor that determines the rate and thus also the extent of N20
decomposition in the reactor.
Emission levels of nitric oxide, however, are thermodynamically limited by
the equilibrium concentration, which increases with temperature (see Fig. 8).
Moreover, as explained in Section 3.2.2, high temperatures favour gas-phase
reaction pathways leading to NO, whereas preferential formation of N20 is
expected at low temperatures. Thus, it is not surprising that declining N20
emission levels have been reported with increasing temperature in nearly all
studies; whenever measured, NO was found to exhibit exactly the opposite
trend (e.g. [126, 135, 136, 151-171]). The observed temperature effect for N20 is
in agreement with modelling performed by Grimsberg and Karlsson [52] (compare Fig. 12) and by Kilpinen and Hupa [53]. It should be pointed out that
temperature distribution along the entire combustor is important, not only bed
temperature.
Johnsson [172] pointed out another effect that may be responsible for the low
NO emissions observed at low temperatures. According to him, it is a high
concentration of CO sustained during low-temperature combustion that leads
to strong enhancement of the (char-catalysed) reduction of NO to N2 (reaction
R45). Similarly, fuel-rich conditions (e.g. in staged combustion) appear to
favour NO removal through this mechanism.
4.1.2 A i r staging and excess air

Air staging is one of the techniques used to reduce NOx levels but its effect
on N20 emissions is unclear. Kramlich et al. [114] reported no change in N20
levels in a laboratory coal-fired furnace in the stoichiometric ratio range
0.4-0.9. Similarly, no clear trend was found in N20 emissions upon staging in
large-scale CFBC's by Hiltunen et al. [157] and in a laboratory-scale pressurised fluidised-bed combustor (PFBC) by Lu et al. [168]. In a 1 MW atmospheric
fluidised-bed combustor (AFBC), however, air staging resulted in a reduction

44

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

of both N20 and NOx levels [159], A similar effect was observed in a laboratoryscale bubbling-bed combustor [158] but N20 emissions were increased and NOx
emissions were reduced upon staging in laboratory-scale [170] and large-scale
[162] PFBC's. Finally, Young et al. [169] found that increasing the primary air
ratio tends to (a) increase N20 emissions at lower temperatures and (b) decrease N20 emissions at higher temperatures (a pilot-scale CFBC).
An increase in excess air was reported to increase NOx but not N20 in
a large-scale AFBC [159], in a PFBC [162] and in an entrained-flow reactor
[152]. In most other studies, an increase in excess air increased N20 emissions
[135, 136, 155 159, 161, 163 165, 169 171] and usually also NOx emissions
[135, 136, 155, 156, 158 160, 162-164, 170, 171]. These observations may be interpreted as a result of enhanced HCN and NH3 oxidation at higher oxygen levels
and are predicted by the modelling done by Grimsberg and Karlsson [52] and by
Kilpinen and Hupa [53] (see Fig. 14).
The discrepancies between the results of different studies are probably due
to the fact that changes in air staging or excess air are difficult to implement
without affecting other important combustor operating characteristics, notably char concentration and temperature distribution along the reactor. This
was demonstrated by mand and Leckner [173] who related N20/NO emission
levels to changes in boiler operation conditions induced by varying the excess
air and fuel-to-air ratios. For example, depending on whether changes in the
excess air ratio were implemented by changing the fuel feed rate or changing
the air flows and fuel feed rate, the NO levels could exhibit opposite trends in
the same CFBC that operated under otherwise identical conditions. Evidence
was also provided for the existence of the fuel-to-air ratio effect independent of
the influence of temperature. The authors concluded that the beneficial effect
of low oxygen concentration on N20 emission was even stronger than that for
NO. Experiments carried out by Aho et al. [153] showed that this could be
explained by enhanced N20 destruction under low-oxygen conditions.

4.1.3 Fuel O,pe

The effect of fuel type is complex and it comprises the following constituents:
(1) elemental composition (in particular fuel-nitrogen content),
(2) volatile-matter content,
(3) composition and content of mineral matter,
(4) chemical nature of bound nitrogen and other elements (functional groups),
(5) physical characteristics of char surface (pore structure).
Some of these properties are lumped into the concept of coal rank and it is
often convenient to relate combustion behaviour of coal to its rank.
Fuel-N conversion to N20 in a laboratory-scale bubbling FBC was found to
increase with coal rank, as shown in Fig. 17. A similar trend, although
somewhat obscured by data scatter, was found for NO, except that the temperature dependence was opposite. These findings can be attributed to the larger
chemical and physical complexity of char surface for low-rank coals, which
makes them more reactive in heterogeneous N20/NO destruction mechanisms.

45

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

50

,10

3O

z"
A,

20

z
10

0
850

950

1050

1150

1250

Temperature [K]

Fig. 17. The effect of temperature and coal rank on fuel-N conversion to N20 in a laboratoryscale AFBC (W6jtowicz et al. [126]). Materials: (+) anthracite (94.1% C, 4.5% VM), ( x )
semianthracite (92.4% C, 11.6% VM), ((~) high-volatile bituminous coal (88.2% C, 31.9%
VM), ( ) high-volatile bituminous coal (81.2% C, 40.4% VM), (C) subbituminous coal (75.0%
C, 53.3% VM), ( # ) lignite (65.9% C, 53.9% VM); carbon and volatile-matter contents on a dry
ash-free basis.

This explanation was further corroborated by burning an activated carbon in


the same system. The fuel-N conversion towards N20 was found to be very low,
as expected. F u r t h e r evidence for the importance of char-surface properties
comes from the study by Pels et al. [103], who combusted chars derived from
different coals in a laboratory FBC and found the same rank dependence as for
parent coals.
The effect of the coal volatile-matter content is also important. It is clear
from Fig. 17 t hat conversion to N20 decreases as the volatile-matter content of
coal increases. This is in agreement with data reported for bubbling beds by
Shimizu et al. [158] and Braun et al. [160], whose comparisons were based on
three and two coals, respectively. A similar dependence of N20 emission on
fuel volatile-matter content was also found in CFBC's [163, 173]. A number of
experimental studies show t hat although NO emission usually decreases with
fuel volatile-matter content in bubbling fluidised-bed combustors (BFBC's), the
trend is reversed in CFBC's [174].
The effect of volatile-matter content on NO emission from BFBC's and
CFBC's can be explained on the basis of qualitative description of combustor
operation. During devolatilisation, a cloud of volatiles surrounding each coal
particle is formed. In a bubbling bed, the volatiles readily enter the bubble

46

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

phase where t h e y form pockets of oxygen-lean gas. Volatiles are s u b s e q u e n t l y


t r a n s p o r t e d to the splash zone immediately above the fluidised bed, and t h e n
into the lower part of the freeboard. There, t h e y form an oxygen-lean environm e n t which is c o n d u c i v e to gas-phase NO reduction. In the case of low-volatile
(i.e. high-rank) materials, this effect is weak, and this leads to h i g h e r NO
emission levels. As discussed by H i r a m a et al. [175], the above m e c h a n i s m is of
no c o n s e q u e n c e in CFBC's, where pockets of volatiles are b r o k e n up and mixed
with oxygen by the vigorous m o t i o n of particles. mand and L e c k n e r [174]
point out t h a t NO r e d u c t i o n in a CFBC is expected to proceed primarily via
a h e t e r o g e n e o u s route. This implies t h a t c o m b u s t i o n of high-volatile coals (i.e.
those h a v i n g low fixed-carbon content) should result in high NO emission
levels due to a lower a m o u n t of c h a r present in the system.
It would seem r e a s o n a b l e to expect a similar m e c h a n i s m for N 2 0 to be valid
as well. Yet no difference in the r e l a t i o n s h i p between N 2 0 emissions and
v o l a t i l e - m a t t e r c o n t e n t has been observed between BFBC's and CFBC's. One
e x p l a n a t i o n would be t h a t in a CFBC, the e n h a n c e m e n t of N20 r e d u c t i o n by
c h a r (low-volatile materials) may be o v e r r i d d e n by N 2 0 p r o d u c t i o n coming as
a result of NO r e d u c t i o n on c h a r surface. Some evidence exists t h a t h i g h e r
a m o u n t s of c h a r found in CFBC cyclone m a t e r i a l seem to c o r r e l a t e with h i g h e r
N20 levels [107]. L a b o r a t o r y experiments show, however, t h a t N 2 0 f o r m a t i o n
resulting from NO r e d u c t i o n on c h a r is negligible [127 131] and thus this
e x p l a n a t i o n must be dismissed. It is more likely t h a t the difference in behavio u r between NO and N20 in both types of boilers comes as a result of different
p r o p o r t i o n s of these species o r i g i n a t i n g from volatile- versus char-N. Indeed,
unlike NO, N 2 0 was r e c e n t l y found to come m a i n l y from volatile-N [103],
which makes the c o r r e l a t i o n between N:O levels and volatile m a t t e r hold for
both BFBC's and CFBC's.
Aho and R a n t a n e n [152] studied c o n v e r s i o n of coal-N to N 2 0 d u r i n g peat
c o m b u s t i o n in an entrained-flow reactor. T h e y found t h a t flue-gas N 2 0 levels
decreased as the O/N ratio increased. A similar c o r r e l a t i o n was also r e p o r t e d
by Lu et al. [168] for c o m b u s t i o n of peat and coals in a PFBC. As l a t e r
d e m o n s t r a t e d by Pels et al. [103], the O/N ratio is related to coal r a n k and thus
the observed trend can be viewed as the effect of coal rank. Similarly, M o r i t o m i
et al. [163] found a r e l a t i o n s h i p between N20 emission from a CFBC and the
fuel ratio, defined as the ratio of fixed c a r b o n to volatile-matter content. Again,
this p r o p e r t y is closely related to coal r a n k and the choice of a coal r a n k index
- - be it c a r b o n or volatile-matter content, O/N, or the fuel r a t i o - - is often
arbitrary.
High coal n i t r o g e n c o n t e n t was found to c o r r e l a t e with high emissions of
both N 2 0 and NO,. [155, 164], which is not a surprising result. To f a c t o r out the
effect of n i t r o g e n c o n t e n t in coal, data are often p r e s e n t e d as a f r a c t i o n a l
c o n v e r s i o n of fuel-N to a product. The role of n i t r o g e n f u n c t i o n a l i t y in coal
(pyrrolic-N versus pyridinic-N) N20 and NO f o r m a t i o n is not clear, however.
Pels et al. [103] r e p o r t e d a c o r r e l a t i o n between n i t r o g e n f u n c t i o n a l i t y and coal
r a n k and related it to N20 emissions from an FBC. The r e l a t i v e a m o u n t of
pyridinic-N in coals was found to increase with coal r a n k and so did N20
emissions from the FBC. The two properties seem to c o r r e l a t e but c a u t i o n

M,A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

47

should be exercised in concluding that N~O is formed from pyridinic forms of


nitrogen in coal. Secondary effects may be responsible for this result, such as
the fact that retention of coal-N in char during devolatilisation also increases
with coal rank. More work is needed in this area.
A less obvious influence on N~O/NO~ emissions is t hat of coal sulphur
content. ~ m a n d et al. [176] showed that firing a high-sulphur coal in a CFBC
results in lower NO and higher N~O levels as compared with combustion of
a similar coal having a low sulphur content. Injection of SO~ into the combustor was used to provide evidence t hat the difference in N~O/NO emissions
between the two coals could be accounted for by the difference in SO~ concentration. The transient response of a combustion system to SO~ injection into the
air inlet is shown in Fig. 18. It is evident t hat an increase in SO~ concent rat i on

2000~--

100

SO~
NO

75

1500

i__

.~1000

500

50
v

25

0
Z

O3
0

200

2OO

t~o

150

tO0

100 ---

o
c.j

=
o
o

50

5O

ff

- N2

20

40

60

... CO

0
0
I

L)

8O

TIME ( r a i n. )
Fig. 18. The effect of SO2 a d d i t i o n to t h e p r i m a r y air d u c t of a CFBC (/~mand et al., [176], by
p e r m i s s i o n of t h e publishers, 1992. ~(~ B u t t e r w o r t h - H e i n e m a n n Ltd.) SO2 level: 1470 ppm
r e l a t e d to t h e t o t a l a m o u n t of air (between time = 20 a n d 55 min.); fuel: Polish high-volatile
b i t u m i n o u s coal (C: 83.8%, H: 4.9%, O: 9.1%, N: 1.4%, S: 0.8% d.a.f.); bed m a t e r i a l : sand, no
l i m e s t o n e added; load: 9.0 MW; bed t e m p e r a t u r e : 1123 K; excess air ratio: 1.21; p r i m a r y - a i r
s t o i c h i o m e t r y : 0.73.

48

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

brings a b o u t a sharp decline in NO and an increase in N20 levels. It was


also shown t h a t the effect of SO2 was identical regardless of w h e t h e r changes
in SO2 c o n c e n t r a t i o n were caused by s u l p h u r c o n t e n t in fuel, SO2 injection or
s u l p h u r c a p t u r e by limestone addition. F u r t h e r evidence of the role played by
SO2 in the N20/NO~ c h e m i s t r y comes from gas-phase c o m b u s t i o n e x p e r i m e n t s
performed by Carlsson and Berge [115, 116] (see Section 3.2.2).
The ash c o n t e n t and its composition m a y also have an influence upon N 2 0
emission, and this occurs m a i n l y t h r o u g h c a t a l y t i c r e a c t i o n s t a k i n g place on
ash components. Lignites, for instance, are k n o w n to c o n t a i n a lot of finely
dispersed calcium which happens to be active in c a t a l y t i c N 2 0 decomposition.
This m a y be one of the reasons why c o m b u s t i o n of lignites has been r e p o r t e d to
result in very low N 2 0 emission [45, 151,126].
Oude Lohuis et al. [155] studied the effect of coal particle size on N 2 0 / N O
emission in a l a b o r a t o r y - s c a l e FBC. T h e y found v e r y little or no difference in
the case of lignite and only a slight increase of N20 levels for a b i t u m i n o u s
coal. The r a n g e of particle size was 53-300 gm. The same but significantly
s t r o n g e r t r e n d was r e p o r t e d by K h a n e t a ] . [171] for a particle size of 1.50 and
4.05 ram.
4.I.4 L i m e s t o n e a d d i t i o n

L i m e s t o n e addition has been d e m o n s t r a t e d to h a v e a beneficial effect on


r e d u c i n g N20 emissions [24, 157, 165, 169, 177 179]. In some cases, however, no
effect was noticed [161, 164, 168, 178, 179] and in two cases an increase in N20
was observed [169, 180]. L o w e r N 2 0 levels were usually offset by an increased
emission of NO [66, 110, 157, 161,164, 176-185] (Fig. 19). In at least two studies,
a good c o r r e l a t i o n was found between the c h a n g e in N O / N 2 0 emissions and
the effective surface area of the added limestone [115, 176].
The effect of limestone on N20 emission levels seems to be complex and is
governed m a i n l y by the c a t a l y t i c scheme (24) (32), gas-phase N20/NOx/SOx

500

500

4O0

40o

300

3O0

o
200

200

100

I
0
1OOO

1100

Temperulure

I
1200

[K}

100

0
1001

~
1100

Temperalure

1 !
1200
[K]

Fig. 19. The effect of limestone addition on N20 and NO emission levels (after Moritomi et
al. [177]). Idemitsu B coal combusted in a CFBC: silica sand (open symbols) and limestone
(solid symbols) used as bed material.

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 -71

49

interactions, and by reactions involving NHz and HCN (e.g. R62). Calciumoxide-catalysed N20 decomposition (reactions 24 and 25) is certainly a decisive
factor in reducing N20 concentration. F u r t h e r lowering of N20 may result
from reduced SO2 levels, as discussed in Section 3.2.2. Yet it has been observed
th at limestone addition usually results in only a moderate decrease in N20
levels, typically by 0 30% [178]. (A nearly 80% decrease in N20 emission at
1123 K which is shown in Fig. 19 is a result of excessively large amounts of CaO
present in the system. The entire bed material was replaced with limestone in
th at experiment, a situation unlikely to occur in industrial practice). This
modest effect of limestone on N20 is surprising in view of highly favourable
kinetics of CaO-catalysed N20 decomposition [113, 132 134, 147]. One reason
for th at may be the fact t hat part of the limestone present in a combustor is
sulphated and thus relatively inactive [178]. In addition, NzO may be formed
via reaction (28) as a by-product of CaO-catalysed NHz oxidation to NO (see
below) [132, 147]. Steady-state formation of N20 has also been observed over
CaO surface in the simultaneous presence of HCN and NO as well as HCN and
02 [147, 186]. A net result of the entire scheme depends on fuel and combustion
conditions but overall reduction in N20 levels is a usual outcome of limestone
addition.
Increased emissions of NO due to limestone injection result from several
effects. First of all, catalytically enhanced oxidation of NHz to NO has been
well documented in the literature [132, 147, 175, 183, 187-189]. Secondly, limestone addition leads to a decrease in SO2 levels which in t urn has been shown
to bring about an increase in NO emission [115, 176]. ,~mand et al. [176] and
Dam-Johansen et al. [190] indicate that CO may also be implicated in the
reaction scheme, perhaps in more than one way (Ca-catalysed CO oxidation
and a gas-phase interaction with SO2). Carbon monoxide is well-known to be
instrumental in NO reduction on char surface [57, 191,192], mainly t hrough
reactions (R40) and (R45). Recent work by Hulgaard [38] also shows that
similar reduction also occurs on limestone, especially when SO2 is present, as
discussed in Section 3.3.2. Finally, increased NO levels resulting from limestone addition may also be partly due to the CaO-catalysed conversion of HCN
to NH3 according to reaction (R62) [66]. An alternative mechanism has also
been proposed [193 195]:
CaO + 2HCN

, CaCN 2 + CO + H2

CaCN2 + H20 + 2H2 +

CO 2

CaO + 2NHa + 2CO

but the validity of either scheme still needs to be demonstrated. The above
HCN/NH3 redistribution would also contribute to a reduction in N20 levels.
In summary, N20 levels decrease, whereas those of NO increase when
limestone is added to the combustion system. The described effect seems to be
stronger for NO than N20.
4.1.5 Injection of an additive (NH3, urea and cyanuric acid)

Ammonia, urea, or cyanuric acid injection is one of the pollution control


measures directed to reduce NOx emissions and this technique is often termed

50

M.A. WOjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

selective non-catalytic reduction (SNCR). In general, SNCR involves NOx


reduction to N2 and H20 according to reaction (R23) and the process is usually
carried out in the temperature window 1173-1373 K. Such a high temperature
is necessary to ensure that NH2 can be formed from NH3.
A similar method of NOx abatement involves ammonia injection into the flue
gas combined with the use of a solid catalyst (typically V205 and TiO2). The
process is known as selective catalytic reduction (SCR), proceeds mainly via
reaction R48, and it is normally utilised in the temperature range 573-673 K.
Low-temperature SCR (operating below 475 K) has recently attracted a lot of
attention. The main advantages offered by this process are dust- and sulphurfree operation as well as the ease of retrofitting at the tail end of the flue-gas
line, just before the stack [196]. A general problem associated with both SNCR
and SCR is emission of excess ammonia into the atmosphere, the so-called
ammonia slip. Ammonia storage, handling and transport may pose additional
difficulties as most countries require special conditions to be met. Therefore,
the use of urea is often preferred, especially in small-scale applications.
A detrimental effect of SNCR upon N20 levels has been reported
[157, 158, 160 162, 170, 197-199], as discussed in detail by Hjalmarsson [200].
The formation of N20 occurs in a relatively narrow temperature range,
1123 1273 K, where NOx reduction is most efficient. In general, the increase in
N20 depends on temperature, the additive used and its feed rate, location of the
injection point, and on the efficiency of NOx reduction. The increase in N20
emissions due to SNCR (AN20) shows an approximate correlation with the
amount of NOx reduced by this process (ANOx). The value of AN20/ANOx has
been reported between 5 and 50% [200]. According to Leckner et al. [201],
however, an increase in N20 can be avoided, even at molar ratios of NH3/NO
as high as three to four.
The main path for the formation of N20 is the reaction of NO with NCO
radicals generated by the decomposition of urea or cyanuric acid in the
combustion chamber (reaction R14 in Table 2). In processes using ammonia,
NO reacts with NHI rather than NCO radicals and the production of N20 is
substantially lower. Caton and Siebers [202] found that ammonia injection
(thermal De-NOx process) produced an order of magnitude less N20 than
injection of cyanuric acid (RAPRENOx process). Cyanuric acid is known to
decompose to form HNCO which enters the HCN oxidation chemistry. The
data by Caton and Siebers are consistent with the discussion of the relative
importance of HCN and NH3 for N20 formation which is presented in Section
3.2.2.
A few results show that it is possible to use urea for NO~ reduction without
a simultaneous increase in N20 emission. There are suggestions that if urea is
decomposed under controlled conditions just before it is injected into the
furnace, the formation of N20 may be prevented [200]. When urea was injected
with over-fire air, N20 formation was found to be negligible [198]. In another
test, natural gas afterburning combined with urea injection resulted in less
N20 formed than during the urea injection alone [203].
The effect of the additive used in the SNCR process was studied by Muzio in
a small pilot-scale combustor [199]. It was found that injection of ammonia

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

51

yielded the lowest N20 emission, 1 8% of ANOx, and the c o r r e s p o n d i n g values


of AN20/ANO.~ for u r e a and c y a n u r i c acid were 7-22% and 15-41%, respectively. Testing was carried out in the t e m p e r a t u r e r a n g e 1100-1370 K and a maximum NOx r e d u c t i o n was observed at 1250 K for a m m o n i a and urea, and at
1370 K for c y a n u r i c acid. The t e m p e r a t u r e at which m a x i m u m N 2 0 was formed
d u r i n g u r e a i n j e c t i o n was also found to be 1250 K. The c o n v e r s i o n of NO to
N20 did not seem to strongly depend on the a m o u n t of additive injected (i.e. on
the N/NOx ratio). N e i t h e r was it found to depend on the initial NOx c o n c e n t r a tion in the r a n g e 300 700 ppm.
Finally, it should be pointed out t h a t the presence of limestone may severely
complicate the c h e m i s t r y of NH3 c o n v e r s i o n in an FBC. D e p e n d i n g on conditions, a m m o n i a may act e i t h e r as a gas-phase N O / N 2 0 p r e c u r s o r (reactions
R1-R8, R36) or as an NO r e d u c i n g a g e n t (e.g. r e a c t i o n s R23, R48 and R47).
I n t r o d u c t i o n of a m m o n i a into a c o m b u s t i o n system o p e r a t i n g with limestone
injection may bring a b o u t changes which are different from the i n t e n d e d ones.
Combined SOx/NO~ control measures k n o w n as E-beam processes m a y r e s u l t
in i n c r e a s e d N 2 0 emission [200]. In E-beam processes, e l e c t r o n beams are used
as an e n e r g y source to excite gas molecules and facilitate SO~/NOx removal. In
one test, the p r o d u c t i o n of N 2 0 was found to be r e l a t i v e l y small, a b o u t 10 ppm
[204].
T h e r e has been c o n c e r n about possible f o r m a t i o n of N20 due to SCR but the
available pilot- and commercial-scale data show t h a t this is not the case
[198, 205, 206]. This is surprising because l a b o r a t o r y studies carried out on
V205 c a t a l y s t s indicate t h a t significant a m o u n t s of N 2 0 can be formed d u r i n g
SCR [207, 208]. L o w - t e m p e r a t u r e SCR has also been r e p o r t e d to r e s u l t in
appreciable release of N20 but new catalysts have been developed which
e n s u r e v i r t u a l l y N20-free NOx removal [209]. According to Topsoe et al. [208],
w a t e r v a p o u r may g o v e r n the distribution of SCR products. In the absence of
w a t e r (most l a b o r a t o r y studies), r e a c t i o n i n t e r m e d i a t e s such as NHxNO undergo d e h y d r o g e n a t i o n to form N20, whereas the presence of w a t e r (industrial
flue gas) promotes d e h y d r a t i o n to N2.
In summary, t e m p e r a t u r e has been identified as the most i m p o r t a n t single
p a r a m e t e r t h a t controls N20 levels in an FBC, high t e m p e r a t u r e leading to
r e d u c e N 2 0 emissions. Very little flexibility is available in using c o m b u s t i o n
t e m p e r a t u r e as a c o n v e n i e n t control p a r a m e t e r to r e d u c e the c o n c e n t r a t i o n of
N 2 0 in the flue gas. A t e m p e r a t u r e typical for fluidised-bed c o m b u s t i o n
( ~ 1123 K) happens to coincide with optimum conditions for SOx removal and
also with r e a s o n a b l y low NO~ levels. The effects of o t h e r p a r a m e t e r s are
summarised in Table 5.
4.2 F u e l - N c o n v e r s i o n to N 2 0 a n d N O

F o r a given coal, the sum of fuel-N c o n v e r s i o n s to N 2 0 and NO in a laboratory-scale FBC was found to be r e m a r k a b l y c o n s t a n t over a r a n g e of tempera t u r e s [126] (Fig. 20). A similar result was i n d e p e n d e n t l y arrived at by G a v i n
and D o r r i n g t o n [210]. The inverse n a t u r e of N 2 0 and NO f o r m a t i o n in the gas
phase, as discussed in Section 3.2.2, would explain this b e h a v i o u r if the

52

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

TABLE 5
The effect of various parameters on N20 and NO emissions in FBC. ( I") Emission increases as
the parameter increases, ( $ ) emission decreases as the parameter increases, and(--)no effect
observed
Parameter

N20 emission

NO emission

Temperature
Coal volatile-matter
content
Coal nitrogen content
Excess air
Air staging
Limestone addition
SNCR
SCR
SO2 level

$
~ (Bubbling FBC)
~ (CFBC)
~
~
~
~
T

T
~ (Bubbling FBC)
T (CFBC)
T
T
T

100

80

60

~)l

900

950

1000

1050

1100

1150

1200

1250

1300

Temperature {K]

Fig. 20 Conversion of coal-N to N 2 0 + N O in a laboratory-scale AFBC. Materials: ( + )


anthracite (94.1% C, 4.5% VM), ( x ) semianthracite (92.4% C, 11.6% VM), () high-volatile
bituminous coal (88.2% C, 31.9% VM), ( 0 ) high-volatile bituminous coal (81.2% C, 40.4%
VM), ( ~ ) subbituminous coal (75.0% C, 53.3% VM), ( ~ ) lignite (65.9% C, 53.9% VM); all coal
characteristics on a d.a.f, basis.
c o n t r i b u t i o n f r o m c h a r - N w e r e n e g l i g i b l e . T h i s p r o v e d n o t to b e t h e c a s e , a t
l e a s t n o t f o r t h e c o a l s a n d c o m b u s t i o n c o n d i t i o n s u s e d t o c o l l e c t t h e d a t a in
F i g . 20 [126, 103]. T h e r e f o r e , a n e x p l a n a t i o n for t h e i n v e r s e r e l a t i o n s h i p bet w e e n c h a r - d e r i v e d N O a n d c h a r - d e r i v e d N 2 0 is s o u g h t . C l a r i f i c a t i o n o f t h i s

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

53

point is important because the trade-off between N20 and NO has also been
found in many large-scale systems (e.g. Braun [211]).
Following the general discussion presented in Section 3.3.1, three heterogeneous N20 formation routes are examined below to see if they may lead to
competitive emissions of N20 and NO.
(1) The oxidation of char-N to N20/NO
02 + (C) + (CN)
, (CO) + (CNO)
N20 formation:
(CN) + (CNO)
2(CNO)

(R30)

, N20 + 2(C)

(R33)

~ (CO) + (C) + N20

(R34)

NO formation:
(CNO)
, NO + (C)

(R31)

(CN) + (CO)

(R32)

, NO + 2(C)

(2) NO + char (or char-N)


*N20
NO + 2(C)
, (CN) + (CO)
NO + (CN)

, N20 + (C)

(3) char-N
,HCN, NH 3
,N20, NO
H2 + (C) + (CN)
, (CH) + (CHN)

(R38a)
(R39)
(R54)

HCN formation:
(CH) + (CN)
, HCN + (C)

(R55)

(CHN)

(R56)

, H C N + (*)

NH3 formation:
H2 + (CHN)
, NH3 + (C)
N20/NO formation:

(R57)

Gas-phase reaction (see Section 3.2.2)

It is apparent from scheme (1) that oxidised surface sites (CNO) may play the
role of a common N20 and NO precursor, in a similar way in which radical
NCO behaves in the gas phase. If reaction (R31) rather than (R32) is the
predominant NO formation pathway, then the NO-versus-N20 trade-off is quite
obvious. Even if reaction (R32) plays a role in the scheme, it should be noted
that the nitrogen-containing site (CN) is also a common substrate in NzO and
NO formation routes (R33) and (R32). The competition of N20 and NO formation routes for the precursor (CNO) can be described in the following way. At
very low temperatures, oxidation of char-N via reaction (R30) is too slow to
produce appreciable amounts of (CNO) and release of N20 and NO does not
occur. At somewhat higher temperatures, N20 and NO formation begins
mainly via reactions (R33) and (R31). NO is expected to be a more abundant
product of (CNO) desorption since N20 formation involves a reaction between
two neighbouring sites. This is consistent with results of packed-bed experiments by De Soete who found that only 1 6% of char-N reacted towards N20

54

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

[45, 65]. At still higher temperatures, the rate of (CN) oxidation to (CNO)
increases and so does the rate of NO production. Formation of N20 is inhibited,
however, due to insufficient inventory of (CN). It is possible that the contribution of reaction (R34) to N20 production is somewhat higher now than at low
temperatures but still insufficient to compete with rapid reaction (R31). The
above scheme accounts for the constancy of N 2 0 + N O levels over a range of
temperatures as well as for decreasing N20 and increasing NO formation rates
with temperature.
It would be difficult to justify the NO-versus-N20 trade-off on the basis of
mechanism (2) but the contribution of this pathway to the overall N20 formation has been found negligible [127-131]. The NO-versus-N20 trade-off in
mechanism (3) is self-evident. Char-N is converted to HCN and NH3 and
N20/NO formation proceeds in the gas phase.
The above discussion shows that in principle it is possible to explain the
NO-versus-N20 trade-off on the basis of both gas-phase and heterogeneous
chemistry. It should be borne in mind, however, that the arguments for the
constancy of NO + N20 emission over a range of temperatures are valid only
for N20/NO formation. In other words, they are strictly correct only if the
N20/NO destruction component is insignificant or invariant in magnitude.
This condition is unlikely to be satisfied in all combustion systems and thus
NO + N20 emissions have often been observed to vary with temperature. Yet
the temperature trade-off between NO and N20 emissions is a common feature
of all combustors. Until recently it has been unclear to what degree each of the
three mechanisms of heterogeneous N20 formation is consequential in operation of large-scale FBC's. In particular, the question of heterogeneous NO
reduction as a source of N20 has attracted a lot of attention. Another important subject is the relative contributions volatile-N and char-N make to overall
N20/NO emissions. The following paragraphs summarise recent advances in
these areas.
Investigation of N20 formation from NO was carried out by ,~mand and
Leckner [173] and Amand et al. [107] in their study of N20/NO emissions from
8 and 12 MW CFBC's. The technique used in the latter work consists in forcing
transient changes in the combustion environment by introducing additional
batches of fuel into the combustor and by stopping the fines and fly-ash
recycle. The issue of relative contributions to N20 emission coming from
volatile-N (HCN) and char-N was addressed in this study. Only low levels of
HCN were detected in the bottom part of a CFBC ( _< 40 ppm) and the concentration of this species was found to rapidly decay along the height of the
combustor. At the same time, N20 concentration increased along the combustor to about 150 ppm in the flue gas. This was interpreted as evidence for the
gas-phase contribution to N20 emission being insignificant. Furthermore, the
work by De Soete [45, 65] was quoted to show that large amounts of N20
released above the devolatilisation zone could not be accounted for by an
approximately 5% char-N conversion to N20. It was concluded that NO reduction on char surface was an important source of N20 in the CFBC.
Moritomi et al. [163] reported results that contradict this conclusion. Injection of NO into a CFBC was found to result in only a small increase in N20

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

55

levels, which means that the contribution of NO reduction to the overall N20
formation is insignificant. In a more recent paper, ~ m a n d et al. [176] reported
results of an experiment similar to the one carried out by Moritomi et al. This
time an appreciable increase in N20 was found upon NO injection into
a 12 MW boiler but gas-phase reactions leading to N20 formation (e.g. R14, RS)
might have been important as well. A coal with a volatile-matter content of
35.7% was used in that study, in contrast to petroleum coke (VM = 5.3%)
combusted by Moritomi et al. In view of the experimental evidence provided by
de Soete [127], Moritomi et al. [163] and other researchers [128-131]. one must
then conclude t hat the amount of N20 resulting from NO reduction on char is
unlikely to be significant.
The argument involving material balance of N20 presented by ~ m a n d et al.
[107] remains generally valid, however, except t hat NO reduction leading to
N20 formation probably occurs on catalytic surfaces rat her than char (e.g.
ash, sulphated limestone; see also Fig. 16). Although no limestone was eraployed in experiments by ,~mand et al. [107], the CFBC operated under fly-ash
recycling conditions. The fact that a sudden discontinuation of fly-ash recycling brought about an increase in NO and decrease in N20 levels renders more
support to the above interpretation.
Pels et al. [103] determined the relative contributions to NaO/NOx emissions
from volatile-N and char-N in a bench-scale bubbling FBC. In this study,
conversion of fuel-N to N20 and NO was examined for a range of coals and
their chars. The contribution of char-N to N20/NO emissions was determined
directly from char combustion experiments, whereas the contribution of volatile-N was calculated by difference. The data were processed assuming that
the conversions of volatile-N and char-N to the products were independent
processes. Elemental analysis of coals and chars as well as thermogravimetric
determination of the amount of volatiles evolved during char preparation were
also used in data evaluation. Chars were prepared by coal pyrolysis in a tube
furnace at temperatures typical for fluidised-bed combustion (1073 1173 K).
The distribution of coal-N between char-N and volatile-N was found to be in
good agreement with the data by Baumann and M611er [98] whose chars were
prepared in a fluidised-bed pyrolyser. The results of the study by Pels et al. are
summarised in Fig. 21. It should be pointed out t hat only net yields of N20, NO
and N2 are considered here, without addressing the issue of relative magnitudes of product formation versus destruction. Examination of data in Fig. 21
leads to the conclusion that in a bubbling FBC, N20 is derived principally from
the volatiles, whereas heterogeneous reactions seem to play an important role
in NO formation. The contribution from char-N to N20/NO emissions increases with coal rank, presumably due to increasingly larger coal-N retention
in char upon pyrolysis (see Section 3.1).
Indirect support for this result comes from the recent experiments by Suzuki
et al. [212, 213] and Fujiwara et al. [166]. In the former work, chars were
prepared at various temperatures and subsequently combusted in two packedbed reactors. Both reactors had the same geometry, except that in one of them
the outlet gas was quenched more rapidly than in the other. Combustion tests
showed that outlet concentrations of NO were similar in both reactors, whereas

56

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


A

[]

N20
fromv~uil~

N2o
from~

[]

NO

~5
7"

from votanle~

NO

from~af

0
65.9

75

81.2

88.2

92.4

94.1

carbon content of coal [% d.a.f.]

Fig. 21. The percentage of original coal-N converted to N20 and NO via volatile-N and
char-N (Pels et al. [103]). Materials: (A) lignite, (B) subbituminous coal; (C) and (D) bituminous coals, (E) semianthracite and (F) anthracite. Experimental: laboratory-scale AFBC,
combustion temperature 1173 K.
N20 levels were lower in the r e a c t o r with a s h o r t e r residence time. F u r t h e r more, char-N c o n v e r s i o n to NO was found to c o r r e l a t e with c h a r p r e p a r a t i o n
t e m p e r a t u r e , whereas the c o n v e r s i o n to N20 exhibited a complex behaviour. It
was concluded t h a t unlike N20, NO f o r m a t i o n was more related to c h a r r a t h e r
t h a n volatile combustion. The d e p e n d e n c e of N 2 0 / N O emissions upon pyrolysis t e m p e r a t u r e was a t t r i b u t e d to the i m p o r t a n c e of the char-N c o n v e r s i o n to
N 2 0 / N O via HCN/NH3 intermediates.
F u j i w a r a et al. [166] carried out bubbling-bed c o m b u s t i o n experiments using
different size of bed particles (sand). T h e y found t h a t NO emission increased as
bed m a t e r i a l size increased, whereas N 2 0 showed the opposite trend. Since the
size of bed particles is related to the distribution of air between the bubble and
dense phases (more air in the dense phase with i n c r e a s i n g particle size), and
since excess air is related to NzO/NOx levels, the a u t h o r s hypothesized t h a t
N20 was formed principally by the volatile c o m b u s t i o n in the bubble phase,
whereas NOx resulted m a i n l y from c h a r c o m b u s t i o n in the emulsion phase.
Thus, different N 2 0 f o r m a t i o n and d e s t r u c t i o n p a t h w a y s seem to be import a n t in bubbling and c i r c u l a t i n g FBC's: the gas-phase r o u t e in the BFBC and
h e t e r o g e n e o u s NO r e d u c t i o n in the CFBC. The following q u a l i t a t i v e description of N20 f o r m a t i o n is c o n s i s t e n t with k n o w n e x p e r i m e n t a l results. As a coal
particle enters an FBC, a plume of volatiles is formed a r o u n d the particle and

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

57

this plume constitutes a reducing environment, a pool of nitrogen-containing


radicals which are instrumental in efficient conversion of NO into N20 and N2.
(The chemistry of these radical reactions, involving mostly NCO and NH~, has
been discussed in Section 3.2.2). In a bubbling FBC, the volatiles enter the
bubble phase in which they are transported to the splash zone above the
fluidised bed. Having reached the splashing zone, N20 can virtually undergo
only gas-phase decomposition due to low concentration of solid particles in the
freeboard. As a consequence, and in view of relatively short residence times
characteristic of most FBC's, appreciable amounts of N20 survive their passage through the freeboard. This effect is further enhanced by the usually
decreasing temperature gradient along the freeboard. Under such conditions,
the amount of N20 formed from the volatiles is by far larger than that formed
from char-N, unless an extremely low-volatile coal is combusted. (This is
actually the case with the a nt hr aci t e in Fig. 21). In addition, N20 formation
from char-N occurs mainly in the dense phase at the bottom part of the
combustor where oxygen concentration is usually low. This may have an
inhibiting effect on char-N conversion to N20.
In a CFBC, on the other hand, char combustion takes place everywhere in
the combustor. Products of gas-phase reactions occurring at the bottom of the
combustor are immediately well mixed by the turbulent motion of solid particles. In this way, N20 that has been formed from volatile-N runs a good
chance of being reduced to N2 by both gas*phase and heterogeneous pathways.
The latter are thought to be of paramount importance in CFBC's due to an
excellent contact between the gas phase and the solids. As a result, N20 that
has been derived from the volatiles may be nearly completely destroyed along
the combustor. This leads to the situation in which a larger proportion of N20
comes from char combustion and/or from NO reduction on catalytic surfaces, if
present. The above description is in general agreement with the discussion of
NO formation in bubbling and circulating FBC's presented by mand and
Leckner [174]. They pointed out that operation of bubbling FBC's is essentially
dominated by gas-phase reactions (volatiles), whereas heterogeneous reactions
play an important role in CFBC's (char). More data are certainly needed to
further substantiate the above-described fate of N20 in bubbling and circulating beds.
4.3

N20

abatement strategies

N20 has emerged as an environmental pollutant only recently and thus no


emission control technologies have yet been developed for this gas. Nevertheless, a few general comments can be made on this subject after N20 formation
and destruction mechanisms have been discussed. First of all, it is clear that no
one-sided approach to N20 abatement is likely to be successful due to a number
of emission trade-offs, notably the one between N20 and NO levels. Secondly,
one should recognise serious constraints on variation of the parameter to
which all emissions are most sensitive, i.e. temperature. For practical purposes, temperature variation in FBC's is limited to the range 1073-1173 K. The
lower bound for temperature is associated with fuel ignition, incomplete

58

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

combustion and poor sulphur capture; the upper one, with high NOx
emission, melting of ash and again inefficient sulphur removal. Three general
approaches to the question of lowering SOx/NOx/N20 emissions can be
considered:
(1) Minimisation of pollutant emissions through improvements in operating
conditions and process control of boilers.
(2) Innovative combustor design to produce low-emission systems.
(3) Sacrificing emissions of one pollutant for the sake of low emissions of the
others, combined with adopting special measures to reduce excessive levels
of the selected pollutant.
The advantage of approach (1) is its relatively low cost and the fact that it
requires only minor modifications in existing plant. Although highly recommended, it is unlikely to bring all emissions from existing plant below acceptable levels. New processes may emerge, however, which would judiciously implement knowledge of pollutant formation/destruction to produce
low-emission combustors (approach 2).
An example of implementation of approach (3) has been recently proposed by
Leckner [75]. It is suggested that in order to reduce N20 emission, the operating temperature of the boiler should be increased to 1173 K, i.e. to its upper
limit. Then, with a limestone addition corresponding to Ca/S = 3, a sulphur
retention of at least 80% can be maintained. A high concentration of limestone
is instrumental in additional lowering of N20 levels, which can be even further
enhanced by gas afterburning. The penalties paid for these modifications are
a greatly increased NO~ levels (temperature+limestone) and a somewhat
higher consumption of limestone. To reduce the final NOx concentration,
ammonia injection is advocated before the cyclone (SNCR). As an additional
measure to bring down the emissions of both N20 and NO_y, a reduction in the
value of the excess air ratio is proposed. It is believed that through an
improved fuel-feed system and process control the excess air ratio of 1.2 can be
maintained, as contrasted with the typical value of 1.3. The latter measure is
an example of approach (1) discussed above.
Another way of applying approach (3) would be to minimise NO~ and SOx
emissions and then to deal with relatively high N20 levels, as suggested by
W6jtowicz et al. [126]. This could be done, for example, by creating a hot zone
in the freeboard or downstream of the combustor (e.g. gas afterburning) or by
utilising catalysis to enhance N20 decomposition.
As far as can be judged, the following N20/NO~ abatement measures may
prove successful in the future:
(1) Gas afterburning
(2) Catalytic additives to enhance N20 decomposition (either within the combustor or as an end-of-pipe solution)
(3) Reduction of NOx through an aftertreatment of flue gas at lower temperatures (catalysis)
Gas afterburning involves introduction of a gaseous fuel (natural gas, methane, etc.) downstream of a coal-fired zone to create an additional flame. Kramlich et al. [48] used a tunnel furnace fired with high-volatile bituminous coal
and demonstrated that gas afterburning resulted in reduction of N20 emission

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

59

from 7 to 3 ppm. G u s t a v s s o n and L e c k n e r [156, 214] successfully applied methane as well as p r o p a n e a f t e r b u r n i n g in a CFBC cyclone. The r e s u l t i n g reduction in N 2 0 levels was a b o u t 40% at an i n j e c t i o n fuel r a t i o of 0.1. The i n j e c t i o n
fuel ratio is defined as the r a t i o of e n e r g y g e n e r a t e d from c o m b u s t i o n of the gas
to the e n e r g y from c o m b u s t i o n of coal. T h e r e was no d e t r i m e n t a l effect on NO
emission observed and a significant decrease in CO levels occurred. It was not
n e c e s s a r y to supply any additional oxygen to the cyclone in order to sustain
the flame. W h e n it was done, however, CO levels dropped even further. Since
a possible release of SO2 in the cyclone due to increased t e m p e r a t u r e was not
investigated, it is still an open question w h e t h e r the cyclone should be used for
a f t e r b u r n i n g , or w h e t h e r an adiabatic a f t e r b u r n i n g c h a m b e r should be employed d o w n s t r e a m of the cyclone [75].
No c a t a l y t i c methods h a v e been applied to N 2 0 removal in fluidised-bed
c o m b u s t i o n as yet. Two classes of such applications can be envisaged: (1)
addition of c a t a l y t i c m a t e r i a l to the fluidised bed, and (2) c a t a l y t i c aftertreatm e n t of flue gases as an "end-of-pipe" solution. The l a t t e r could be conven i e n t l y realised in a m o n o l i t h i c r e a c t o r and possibly combined with gasa f t e r b u r n i n g . The d i s a d v a n t a g e of the "end-of-pipe" a p p r o a c h is its cost and
difficulties with retrofitting. This is also t r u e a b o u t applying the same idea to
NOx removal from flue gases. An additional c o m p l i c a t i o n is associated with the
possibility of N 2 0 p r o d u c t i o n d u r i n g l o w - t e m p e r a t u r e r e d u c t i o n of NO.
It is expected t h a t i m p l e m e n t a t i o n of the above-described techniques, separ a t e l y or in combination, will result in successful i n c o r p o r a t i o n of N 2 0 control
into FBC technology. As a consequence, fluidised-bed c o m b u s t i o n will emerge
as a t r u l y clean and efficient m e t h o d of power generation.

5. CONCLUDING REMARKS
N i t r o u s oxide is a strong, long-lived absorber of infrared r a d i a t i o n and as
such c o n t r i b u t e s to e n h a n c e m e n t of the g r e e n h o u s e effect. It also acts as
a nitric oxide p r e c u r s o r in the s t r a t o s p h e r e and in this way plays an
i m p o r t a n t role in depletion of the ozone layer. The c o n t r i b u t i o n of N 2 0 to
the g r e e n h o u s e effect has been rising steadily and was about 6% in the
1980's. D i n i t r o g e n oxide raises e n v i r o n m e n t a l c o n c e r n s due to (a) its long
a t m o s p h e r i c lifetime, and (b) the possibility of a massive increase in its
emission upon switching to fluidised-bed combustion. F u r t h e r a u g m e n t a t i o n
of its presence in the a t m o s p h e r e may come from the i n c r e a s i n g use of
c a t a l y t i c c o n v e r t o r s in cars and from some NOx control technologies, such
as u r e a injection. In view of c u r r e n t u n c e r t a i n t i e s a b o u t e n v i r o n m e n t a l
response to global w a r m i n g and the lack of u n e q u i v o c a l evidence for the
cause effect r e l a t i o n s h i p between the g r e e n h o u s e effect and the observed
t e m p e r a t u r e increase over several decades, a cautious but prompt a p p r o a c h
to c u r b i n g N20 emissions is advocated. It is suggested t h a t (1) c o m b u s t i o n
efficiency should be increased by i m p r o v e m e n t s in plant operation, (2) res e a r c h efforts should be directed into gaining a b e t t e r u n d e r s t a n d i n g of the
c h e m i s t r y u n d e r l y i n g N20 emissions from c o m b u s t i o n sources, and (3)

60

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

knowledge resulting from research should be applied to implement modifications to FBC operating r6gimes. Such modifications are expected to result in
a reduction of N20 levels without sacrificing the relatively low emissions of
SOx and NOx.
Nitrogen is present in coal in the form of heterocyclic aromatic structures
(mostly of the pyrrolic and pyridinic type). They exhibit relatively high
thermal stability and, as a consequence, coal devolatilisation under
fluidised-bed conditions results in most of coal-N being retained in the char
(typically more than 60%). HCN appears to be a primary nitrogen-containing product of heterocyclic ring rupture. Subsequently HCN undergoes
hydrogenation to form NH3. The composition and structural characteristics
of tar closely resemble those of the parent coal.
Thermodynamic calculations show that the concentration of N20 in FBC
flue gases is several orders of magnitude larger than its equilibrium value.
This is not true about NO emission, which is comparable to the equilibrium
concentration. Since reaction rates responsible for N20 destruction are
relatively high at temperatures prevailing in the FBC, it appears that N20
formation in a combustor is somewhat delayed and occurs too late for
the destructive mechanisms to successfully reduce N20 levels. Clearly,
any measure directed towards increasing gas residence time in a combustor and/or inducing early N20 release is expected to have a beneficial
effect.
Several observations support the interpretation presented above: (a) rising N20 concentration profiles often reported for FBC's, whereas decaying
patterns prevail in flames; (b) sensitivity of N20 levels to residence time in
a combustor; and (c) lower concentrations of N20 reported for "hot" freeboards in AFBC's (increased residence time at elevated temperature).
Another way of bringing N20 concentrations closer to their equilibrium
values is catalytic enhancement of N20 destruction pathways. More work
needs to be done to explore this option.
Strong evidence has been produced that under FBC conditions, N20 is
formed mainly from cyano species released as volatile-N, whereas NH3-based
compounds tend to react towards NO.
Gas-phase formation of N20 is competitive with respect to NO formation. As
temperature decreases, more N20 is formed at the expense of NO.
Only up to 10% of char-N has been found to form N20. N20/NO destruction
mechanisms on char surface have been demonstrated to be important under
FBC conditions, especially in the presence of CO. Laboratory studies show
that NO reduction on char surface is not a significant source of N20.
The role played by ash, limestone and by other solids of potentially catalytic
properties in the overall chemistry of nitrogenous species is complex and far
from being understood. A lot of progress has been made on solid-catalysed
N20 decomposition but little is known about, for example, solid-catalysed
HCN/NH3 oxidation to N20/NO~, or about a possible solid-catalysed conversion of HCN into NH3 and vice versa. Recent work shows that many
CaO-catalysed reactions occur within the temperature range relevant to
fluidised-bed combustion. Calcined limestone has been found to strongly

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

61

catalyse N20 decomposition to N2. This reaction is not adversely affected by


the presence of other gases (02, NO, NH3 and HCN). Nevertheless, reduction in N20 levels caused by limestone injection into an FBC is r a t h e r
moderate. This may be due to (1) a relative inactivity of sulphated limestone
towards N20 decomposition; and (2) the possibility of N20 formation as
a result of NO reduction catalysed by this material.
Temperature has been identified as the most important single parameter that
controls N20 levels is an FBC, high temperature leading to reduced N20
emissions. Little flexibility is available in using combustion temperature as
a convenient control parameter to reduce the concentration of N20 in flue
gas. A combustion temperature typical for FBC operation ( ~ 1100 K) happens to coincide with optimum conditions for SO~ removal, and also with
reasonably low NO~ levels.
Interactions between sulphur- and nitrogen-containing species in a combustion system are an intriguing question, especially if limestone is involved. In
the gas-phase, the result of reduced SO2 levels is to decrease the net N20
formation and to cause an increase in NO emission. Limestone also exhibits
a catalytic effect leading to N20 decomposition as well as increased NO
production. For both N20 and NO, the effect of heterogeneous reactions
associated with the presence of limestone is reinforced by the homogeneous
chemistry related to reduced SO2 levels. In this way, N20 levels decrease,
whereas those of NO increase when limestone is added to the combustion
system. The described effect is stronger for NO than N20. Some NOx control
technologies, such as SNCR, have been found to result in increased levels
of N20. It should be recognised that a comprehensive approach to combustor
design (with focus on SO~ NO~ N20 interactions) is very much needed
to minimise negative interactions among different pollution control
measures.
A trade-off between N20 and NO emissions has been found to be a striking
feature of nearly all combustion systems. Low emissions of NO at low
temperatures are accompanied by high emissions of N20, with a reverse
behaviour at high temperatures. It is possible to explain this qualitatively on
the basis of cur r ent knowledge of N20/NO chemistry.
Different N20 formation and destruction pathways seem to be important in
bubbling and circulating FBC's. There are some indications that heterogeneous NO reduction may be a predominant source of N20 in a CFBC, whereas
N20 is mainly derived from volatile-N in a bubbling FBC. This may be the
result of gas-phase reactions playing an important role in bubbling. FBC's,
whereas operation of CFBC's is usually dominated by heterogeneous reactions.
As far as can be judged, the following N20/NOx abatement measures may
prove successful in the future:
(1) Gas afterburning
(2) Catalytic additives to enhance N20 decomposition (either within the
combustor or as an end-of-pipe solution)
(3) Reduction of NO~ through an aft ert reat m ent of flue gas at lower temperatures (catalysis)

62

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

ACKNOWLEDGEMENTS
F i n a n c i a l s u p p o r t for t h i s w o r k w a s p r o v i d e d b y t h e C o m m i s s i o n of t h e
E u r o p e a n C o m m u n i t i e s t h r o u g h g r a n t J O U F 0047-C(SMA). Prof. Bo L e c k n e r
of C h a l m e r s U n i v e r s i t y , D r H e r b e r t B a u m a n n of D M T a n d M r N i k l a s B e r g e of
S t u d s v i k E n e r g y m a d e r e s u l t s of t h e i r u n p u b l i s h e d w o r k a v a i l a b l e to us.
P r o f e s s o r L e c k n e r also r e a d t h e m a n u s c r i p t a t t h e f i n a l s t a g e of its p r e p a r a t i o n
a n d g a v e h e l p f u l r e m a r k s . All t h e s e c o n t r i b u t i o n s a r e g r a t e f u l l y a c k n o w ledged.

REFERENCES
1 Houghton, J.T., Jenkins, G.J. and Ephraums, J.J. (Eds.), 1990. Climate change, The
IPCC Scientific Assessment. Press Syndicate of the University of Cambridge, London.
2 Sinclair, J., 1990. Ozone loss will hit health and food, says UN study. New Sci.,
125(1702): 27~
3 Sloss, L.L., 1991. NOx emissions from coal combustion. Report EACR/36, IEA Coal
Research, London.
4 Rodhe, H. and Johansson, C., 1989. Impact of anthropogenic activities on the global
N2 cycle. Report IVL3568, IVL Swedish Environmental Research Institute, Stockholm,
pp. 1 9.
5 Crutzen, P.J., 1983. Atmospheric interactions Homogeneous gas reactions of C, N,
and S containing compounds. In: B. Bolin and R.B. Cook (Eds.), The major Biogeochemical Cycles and Their Interactions. Wiley, New York, pp. 67 112,
6 Seitz, F. (Ed.), 1989. Scientific perspectives on the greenhouse problem. Marshall
Institute, Washington, DC.
7 Hammer, C.U., Clausen, H.B. and Dansgaard, W., 1980. Greenland ice sheet evidence of
postglacial volcanism and its climatic impact. Nature, 288:230 235.
8 Porter, S.C., 1987. Pattern and forcing of the northern hemisphere glacier variations
during the last millennium. Quart. Res., 26:27 48.
9 Bolin, B., D66s, B.R., J~ger, J. and Warrick, R.A. (Eds.), 1986. The Greenhouse Effect,
Climatic Change and Ecosystems. SCOPE 29, Wiley, Chichester, UK.
10 Khalil, M.A.K. and Rasmussen, R.A., 1988. Nitrous oxide: trends in global mass balance
over the last 3000 years. Ann. Glaciol., 10:73 79.
11 Pearman, G.I., Etheridge, D.M. and De Silva, F., 1986. Evidence of changing concentrations of atmospheric CO> N2 and CH4 from air bubbles in Antarctic ice. Nature
320:248 250.
12 Stauffer, B.R. and Neftel, A., 1988. What have we learned from the ice cores about the
atmospheric changes in the concentrations of nitrous oxide, hydrogen peroxide, and
other trace species. In: F.S. Rowland and I.S.A. Isaksen (Eds.), The Changing Atmosphere; Report from the Dahlem workshop on the changing atmosphere. Wiley, Chichester, UK, pp. 63 77.
13 Etheridge, D.M., Pearman, G.I. and De Silva, F., 1988. Atmospheric trace-gas variations
as revealed by air trapped in an ice core from Law Dome, Antarctica. Ann. Glaciol.,
10:28 33.
14 Zardini, D., Raynaud, D., Scharffe, D. and Seiler, W., 1989. N20 measurements of air
extracted from Antarctic ice cores: Implications on atmospheric N20 back to the last
glacial-interglacial transition. J. Atm. Chem., 8:189 201.
15 Elkins, J.W. and Rossen, R., 1989. Summary Report 1988: Geophysical monitoring for
climatic change. NOAA ERL, Boulder, CO.
16 Warr, K., 1990. Ozone: The burden of proof. New Scientist, 125: 36-40.

M.A. WSjtowicz et al./Fuel Processing Technol. 34 (1993) 1-71

63

17 Anderson, I.C. and Levine, J.S., 1987. Simultaneous field measurements of biogenic
emissions of nitric oxide and nitrous oxide. J. Geophys. Res., 92:965 976.
18 Levine, J.S., 1992. The global atmospheric budget of nitrous oxide. A keynote lecture
Proc. 5th Int. Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3,
1992, Tsukuba, Japan, p. KL-I-1.
19 Rodhe, H., 1990. A comparison of the contribution of various gases to the greenhouse
effect. Science, 248:1217 1219.
20 Elkins, J.W., Wafsy, S.C., McElroy, M.B., Kolb, C.E. and Kaplan, W.A., 1978. Aquatic
sources and sinks for nitrous oxide. Nature, 275: 602-606.
21 Blackmer, A.M. and Bremner, 1976. Potential of soil as a sink for atmospheric nitrous
oxide. Geophys. Res. Lett., 3:739 742.
22 Muzio, L.J. and Kramlich, J.C., 1988. An artifact in the measurement of N20 from
combustion sources. Geophys. Res. Lett., 15(12): 1369-1372.
23 De Soete, G.G., 1990. Re-evaluation of N20 emissions from fossil fuel combustion.
LNETI/EPA/IFP European workshop on the emission of nitrous oxide. LNETI, Lisbon,
Portugal, pp. 41 45.
24 Botting, A.J., Gavin, D.G. and Hughes, I.S.C., 1991. Emissions of nitrous oxide from
coal-fired fluidised bed boilers. Proc. 5th Int. FBC Conf., 10 11 Dec., London, Institute
of Energy, London, p. 239.
25 Thiemens, M.H. and Trogler, W.C., 1991. Nylon production: An unknown source of
atmospheric nitrous oxide. Science, 251:932 934.
26 Elkins, J., 1991. Current uncertainties in the global atmospheric N20 budget.
LNETI/EPA/IFP European workshop on the emission of nitrous oxide. LNETI, Lisbon,
Portugal, pp. 25 33.
27 Lobert, J.M., Sharffe, D.H., Hao, W.M. and Crutzen, P.J., 1990. Biomass burning as
a source of atmospheric N20 and NOx. LNETI/EPA/IFP European workshop on the
emission of nitrous oxide. LNETI, Lisbon, Portugal, pp. 35 40.
28 Ryan, J.V. and Srivastava, R.K., 1989. EPA/IFP European workshop on the emission of
nitrous oxide from fossil fuel combustion. EPA-600/9-89-089, U.S. Environmental Protection Agency, Research Triangle Park, NC.
29 Conrad, R., Seller, W. and Bunse, G., 1983. Factors influencing the loss of fertilizer
nitrogen into the atmosphere as N20. J. Geophys. Res., 88:6709 6718.
30 Ronen, D., Mordeckai, M. and Almon, E., 1988. Contaminated aquifers are a forgotten
component of the global N20 budget. Nature, 335:57 59.
31 Matson, P.A. and Vitousek, P.M., 1987. Cross-system comparisons of soil nitrogen
transformations and nitrous oxide flux in tropical forest ecosystems. Global Biogeochem. Cycles, 1:163 170.
32 Luizao, F., Matson, P., Livingston, G., Luizao, R. and Vitousek, P., 1989. Nitrous oxide
flux following tropical land clearing. Global Biochem. Cycles, 3:281 285.
33 Schmidt, J., Seller, W. and Conrad, R., 1988. Emission of nitrous oxide from temperate
forest soils into the atmosphere. J. Atm. Chem., 6:95 115.
34 Bowden, R.D., Steudler, P.A., Melillo, J.M. and Aber, J.D., 1990. Annual nitrous oxide
fluxes from temperate forest soils in the Northeastern United States. J. Geophys. Res.,
95(D9): 13997 14005.
35 Crutzen, P.J., Delany, A.C., Greenberg, J., Haagenson, P., Heidt, L., Lueb, R., Pollock,
W., Seller, W., Wartburg, A. and Zimmerman, P., 1985. Tropospheric chemical composition measurements in Brazil during the dry season. J. Arm. Chem., 2:233 256.
36 Corer III, W.R., Levine, J.S., Winstead, E.L. and Stocks, B.J., 1991. New estimates of
nitrous oxide emissions from biomass burning. Nature, 349 (6311): 689-691.
37 Reimer, R.A., Parrett, R.A. and Slaten, C.S., 1992. Abatement of N20 emissions produced in adipic acid manufacture. Proc. 5th Int. Workshop on Nitrous Oxide Emissions
(NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 10-1-1.
38 Hulgaard, T., 1991. Nitrous oxide from combustion. Ph.D. thesis. Department of Chemical Engineering, Technical University of Denmark, Lyngby, Denmark.
39 Andersson, S., Br~innstr6m-Norberg, B.-M. and Hanell, B., 1989. Nitrous oxide emissions from different combustion sources. Report No. U(V) 1989/31, Vattenfall, Sweden.

64

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

40 Andersson, S., ~mand, L.-E. and Leckner, B., 1988. N20 emissions from fluidized bed
combustion. Paper presented at the IEA AFBC technical meeting, Amsterdam, The
Netherlands, 18 20 Nov. 1988.
41 Davis, W.K., 1990. Energy and the environment. Chem. Eng. Prog., 86(7):64 67.
42 EPA/IFP European workshop on N20 emissions, 1988. Rueil-Malmaison, France, 1 2
June.
43 Roby, R.J. and Bowman, C.T., 1987. Formation of N20 in laminar, premixed, fuel-rich
flames. Combust. Flame, 70: 119-123.
44 Houser, T.3., McCarville, M.E. and Zhou-Ying, G., 1988. Nitric oxide formation from
fuel-nitrogen model compound combustion. Fuel, 67:642 650.
45 De Soete, G.G., 1989.9th Members Conf. of IFRF, Noordwijkerhout, The Netherlands,
24 26 May.
46 De Soete, G.G., 1989. M6canisme de formation et de destruction des oxydes d'azote
dans la combustion (Mechanisms of formation and decomposition of nitrogen
oxides in combustion). Rev. G6n. Therm. Fr., No. 330 331 (June July) 28:353 373 (in
French).
47 De Soete, G.G., 1989. Cin6tique de la formation de pollutants gazeaux dus fi la combustion. Un example type: l'Oxyde et le protoxyde d'azote (Kinetics of gaseous-pollutant
formation in combustion. A typical example: nitric and nitrous oxides). Pollution
Atmosph6rique, April June, 122:176-183 (in French).
48 Kramlich, J.C., Cole, J,A., McCarthy, J.M., Lanier, W.S. and McSorley, J.A., 1989.
Mechanisms of nitrous oxide formation in coal flames. Combust. Flame, 77: 375-384.
49 Martin, R.J. and Brown, N.J., 1990. Nitrous oxide formation and destruction in lean,
premixed combustion. Combust. Flame, 80:238 255.
50 Glarborg, P., Miller, J.A. and Kee, R.J., 1986. Kinetic modelling and sensitivity analysis
of nitrogen formation in well-stirred reactors. Combust. Flame, 65:177 202.
51 Grimsberg, M., 1990. Formation of nitrogen oxides during combustion. Licentiate
Thesis~ Lund University, Sweden.
52 Grimsberg, M. and Karlsson, H.T., 1990. Homogeneous formation and decomposition of
nitrogen oxides during combustion. Paper presented at the Finnish-Swedish Flame
Days, Turku/Abo, Finland, 4 5 September, 1990.
53 Kilpinen, P. and Hupa, M., 1991. Homogeneous N20 chemistry at fluidized bed combustion conditions: a kinetic modeling study. Combust. Flame, 85:94 104.
54 Martin, R.J. and Brown, N.J., 1990. Analysis and modeling of nitrous oxide chemistry in
lean, premixed combustion. Combust. Flame, 82:312 333.
55 Hulgaard, T., Glarborg, P. and Dam-Johansen, K., 1991. Homogeneous formation and
destruction of N20 at fluidized bed combustion conditions. In: E.J. Anthony (Ed.), Proc,
l l t h Int. Conf. on FBC, ASME, New York, pp. 991 998.
56 Miller, J.A. and Bowman, C.T., 1989. Mechanism and modeling of nitrogen chemistry in
combustion. Prog. Energy Combust. Sci., 15:287 338.
57 B/itz, P., Ehbrecht, J., Hack, W., Rouveirolles, P. and Wagner, H.G., 1988. Detailed
kinetic studies on elementary NH2-reactions. 21st Syrup. (Int.) on Combustion. The
Combustion Institute, Philadelphia, PA, pp. 1107 1115.
58 Davidson, D.F., Kohse-Hoinghaus, K., Chang, A.Y. and Hanson, R.K., 1990. Int. J.
Chem. Kin., 22: 513.
59 Davidson, D.F., and Hanson, R.K., 1990. Int. J. Chem. Kin., 22: 843.
60 Mertens, J.D., Chang, A.Y., Hanson, R.K., Bowman, C.T. and Masten, D.A., 1989.
A shock tube study of the reactions of NH with NO, 02, and O. Paper No. WSS/CI-89-96,
presented at the Western States Section/The Combustion Institute meeting, Livermore, CA.
61 Davidson, D.F., Dean, A.M., DiRosa, M.D. and Hanson, R.K., 1990. (Quoted in ref. [38]).
62 Tully, F.P., Perry, R.A., Thorne, L.R. and Allensdorf, M.D., 1988. Free-radical oxidation
of isocyanic acid. 22nd Syrup. (Int.) on Combustion. The Combustion Institute, Phildelphia, PA, pp. 1101 1106.
63 Hanson, R.K. and Salimian, S., 1984. In: W.C. Gardiner, Jr. (Ed.), Combustion chemistry. Springer-Verlag, New York, p. 361.

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


64

65

66

67
68

69

70
71
72

73

74

75
76
77
78
79
80

81
82

83

84
85

65

Loirat, H., Caralp, F., Destriau, M. and Lesclaux, R., 1987. Oxidation of CO by N20
between 1076 and 1228 K: Determination of the rate constant of the exchange reaction.
J. Phys. Chem., 91: 6538.
De Soete, G.G., 1990. Heterogeneous N20 and NO formation from bound nitrogen atoms
during coal char combustion. 23rd Symp. (Int.) on combustion. The Combustion Institute, Philadelphia, PA, pp. 1257 1264.
Gavin, D.G. and Dorrington, M.A., 1991. Basic studies of NOx formation and control.
Final report on ECSC project No. 7220-EA 819. British Coal Corporation, Coal Research Establishment, London, June 1991.
Johnsson, J.E., 1990. Kinetics of heterogeneous NOx reactions at FBC conditions. Dept.
Chem. Eng., Technical Univ. Denmark, Lyngby, Denmark, CHEC report No. 9003.
Johnsson, J.E., 1991. Nitrous oxide formation and destruction in fluidized bed combust i o n - - A literature review of kinetics. Presented at the 23rd IEA AFBC Meeting in
Firenze, Italy, 8 November, 1991. CHEC report No. 9103.
De Soete, G.G., 1989. Formation of nitrous oxide from NO and SO2 during solid fuel
combustion. Proc. 1989 Joint EPA/EPRI Symp. on stationary combustion NO~ control,
6 9 March, San Francisco, CA.
De Soete, G.G., 1988. EPA/IFP European Workshop on the emission of nitrous oxide
from fossil fuel combustion. Rueil-Malmaison, France, June 1 2, 1988.
Santala, P., Iisa, K. and Hupa, M., 1990. Catalytic destruction of N20 in a fixed bed
laboratory reactor. Report 90--6, ,~bo Akademi, Finland. (Quoted in ref. [68].)
Furusawa, T., Honda, T., Takano, J. and Kunii, D., 1978. Nitric oxide reduction in an
experimental fluidized-bed coal combustor. In: J.F. Davidson and D.L. Keairns (Eds.),
Fluidization; Proc. 2nd Engineering Foundation Conf. Cambridge University Press,
London, pp. 314 319.
Pereira, F.J., Beer, J.M., Gibbs, B. and Hedley, A.B., 1974. NO~ emissions from fluidizedbed coal combustors. 15th Symp. (Int.) on combustion. The Combustion Institute,
Philadelphia, PA, pp. 1149 1156.
Jonke, A.A., 1969. Reduction of atmospheric pollution by the application of fiuidized
bed combustion. Argonne National Lab., Publication no. ANL/ES-CEN 1001 to 1004,
Argonne, IL.
Leckner, B., 1992. Optimization of emissions from fluidized bed boilers. Int. J. Energy
Res., 16:351 363.
J/intgen, H. and Van Heek, K.H., 1979. An update of German non-isothermal coal
pyrolysis work. Fuel Processing Technol., 2:261 293.
Howard, J.B., 1981. In: M.A. Elliott (Ed.), Chemistry of coal utilization, 2nd Suppl. Vol.
Wiley, New York, pp. 665 784.
Gavalas, G.R., 1982. Coal Pyrolysis. Coal Science and Technology 4. Elsevier,
Amsterdam.
Berkowitz, N., 1985. The Chemistry of Coal. Coal Science and Technology 7. Elsevier,
Amsterdam.
Suuberg, E.M., 1985. Mass transfer effects in pyrolysis of coals: A review of experimental evidence and models. In: R.H. Schlosberg (Ed.), Chemistry in coal conversion.
Plenum Press, New York, pp. 67 119.
Solomon, P.R. and Hamblen, D.G., 1985. Pyrolysis. In: R.H. Schlosberg (Ed.), Chemistry
in Coal Conversion. Plenum Press, New York, pp. 121 251.
Tromp, P.J.J. and Moulijn, J.A., 1988. Slow and rapid pyrolysis of coal. In: Y. Yfiriim
(Ed.), New Trends in Coal Science. Nato ASI Series, Kluwer Academic, Dordrecht,
pp. 305 338.
Pohl, J.H. and Sarofim, A.F., 1976. Devolatilization and oxidation of coal nitrogen.
16th Symp. (Int.) on combustion. The Combustion Institute, Philadelphia, PA
pp. 491 501.
Kirner, W.R., 1945. The occurrence of nitrogen in coal. In: H.H. Lowry (Ed.), Chemistry
of Coal Utilization. Wiley, New York, pp. 450 484.
Beet, A.E., 1940. A further study of the Kjeldahl process. Pyridine Carboxylic acids
from coal. Fuel, 19: 108.

66
86
87
88

89

90
91

92
93
94
95
96
97
98

99

100

101
102

103
104

105

106

107
108
109
110

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


Deal. V.Z., Weiss, F.T. and White, T.T., 1953. Determination of Basic Nitrogen in Oils.
Anal. Chem., 25: 426.
Montgomery, R.S. and Holly, E.D., 1957. Structures of the acids obtained by oxidation
of bituminous coal Thermal decarboxylation of the copper salts. Fuel, 36:493 500.
Davies, C. and Lawson, G.J., 1967. Chemical constitution of coal; XIV The products of
oxidation of humic acid with three parts of potassium permanganate at 30'C. Fuel,
46:127 136.
Entel, J., 1958. Nuclear Structure of the Water-soluble Polycarboxylic Acids from the
Oxidation of Bituminous Coals: The Decarboxylation Reaction. J. Am. Chem. Soc.,
77:611.
Hayatsu, R., Scott, R.G., Moore, L.P. and Studier, M.H., 1975. Aromatic units in coal.
Nature, 257:378 380.
King, S.B., Brandenburg, C.F. and Lanum, W.J., 1975. Characterization of nitrogen
compounds in tar produced from underground coal gasification. Am. Chem. Soc., Div.
Fuel Chem., Prepr., 20(2): 131.
Hauck, R.D., 1975. Am. Chem. Soc., Div. Fuel Chem., Prepr., 20(2): 85.
Jones, R.B., McCourt, C.B. and Swift, P., 1981. Proc. 1981 Int. Conf. Coal Sci., Dfisseldorf. Verlag Glfickauf. Essen, Germany, p. 657.
Perry, D.L. and Grint, A., 1983. Application of XPS to coal characterization. Fuel,
62:1024 1033.
Wallace, S., Bartle, K.D. and Perry, D.L., 1989. Quantification of nitrogen functional
groups in coal and coal derived products. Fuel, 68:1450 1455.
Burchill, P. and Welch, L.S., 1989. Variation of nitrogen content and functionality with
rank for some UK bituminous coals. Fuel, 68:100 104.
Solomon, P.R. and Colket, M.B., 1978. Evolution of fuel nitrogen in coat devolatilization. Fuel, 57:749 755.
Baumann, H. and M611er, P., 1991. Pyrolysis of hard coals under fluidised bed combustor conditions---Distribution of nitrogen compounds on volatiles and residual char.
Erd6l, Erdags u. Kohle 44(1): 29 33.
Blair, D.W., Wendt, J.O.L. and Bartok, W., 1976. Evolution of nitrogen and other
species during controlled pyrolysis of coal. 16th Symp. (Int.) on combustion. The
Combustion Institute, Philadelphia, PA, pp. 475 489.
Kobayashi, H., 1976. Devolatilization of pulverized coal at high temperatures. Ph.D.
thesis, Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA.
Suuberg, E.M., Peters, W.A. and Howard, J.B., 1978. Product composition and kinetics
of lignite pyrolysis. Ind. Eng. Chem., Proces Des. Dev., 17: 37.
W6jtowicz, M.A., Pels, J.R., van Langeveld, A.D., Sloof, W.G. and Moulijn, J.A., 1992.
Nitrogen functionality in coals and coal chars, In: Proc. Int. Conf. Carbon, AKK DKG,
Essen, Germany, 22 -26 June, 1992, pp. 52 54.
Pels, J.R., Wojtowicz, M.A. and Moulijn, J.A., 1992. Rank dependence of N20 emission
in fluidised-bed combustion of coal. Fuel, 72:373 379.
Bose, A.C., Dannecker, K.M. and Wendt, J.O.L., 1988. Coal composition effects on
mechanisms governing the destruction of NO and other nitrogenous species during
fuel-rich combustion. Energy and Fuels, 2:301 308.
Ghani, M.U. and Wendt, J.O.L., 1990. Early evolution of coal nitrogen in opposed flow
combustion configurations. 23rd Symp. (Int.) on combustion, The Combustion Institute,
Philadelphia, PA, pp. 1281 1288.
Chen, S.L., Heap, M.P., Pershing, D.W. and Martin, G.B., 1982. Influence of coal
composition on the fate of volatile and char nitrogen during combustion. 19th Symp.
(Int.) on combustion, The Combustion Institute, Philadelphia, PA, pp. 1271 1280.
,~mand, L.-E., Leckner, B. and Andersson, S., 1991. Formation of N20 in circulating
fluidized bed boilers. Energy and Fuels, 5:815 823.
Suuberg, E.M., 1977. Rapid pyrolysis and hydropyrolysis of coal. D.Sc. thesis, Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA.
Dry. R.J. and La Nauze, R.D., 1990. Combustion in fluidized beds. Chem. Eng. Prog.,
86(7): 31 42.
Leckner, B. and Amand, L.-E., 1987. Emissions from a circulating and a stationary

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

111
112
113

114
115

116
117

118
119
120
121

122

123
124
125

126

127

128
129

130

131

132

67

fluidized bed boiler: A comparison. In: J.P. Mustonen (Ed.), Proc. 9th Int. Conf. on FBC.
Am. Soc. Mech. Eng., New York, pp. 891 897.
Hayhurst, A.N. and Vince, I.M., 1980. Nitric oxide formation from nitrogen fames. The
importance of "prompt" nitric oxide. Prog. Energy Combust. Sci., 6: 35.
Clarke, A.G. and Williams, A., 1991. The formation and control of NO~ emissions. Chem.
Ind., No. 24: 917-920.
Khan, T., Lee, Y.Y. and Young, L., 1991. Heterogeneous decomposition of nitrous oxide
in the operating temperature range of circulating fluidized bed combustors. EPRI/EPA
Joint symp. on stationary combustion NOx control, Washington, DC, March 25-28.
Kramlich, J.C., Nihart, R.K., Chen, S.L., Pershing, D.W. and Heap, M.P., 1982. Behaviour of N20 in staged pulverized coal combustion. Combust. Flame, 48: 101-104.
Carlsson, M., 1991. Effekt av kalktillsats av kalksten respektive SO2 pfi NO och N20emissioner frfin en gaseldad bubblande bfidd i laboratorieskala. (The effect of calciumoxide addition on the emission of SO:, NO or N20 from a laboratory-scale fluidised
bed). Arbetsrapport-Technical Note, Studsvik UP-91/41, Sweden (in Swedish).
Carlsson, M. and Berge, N., 1992. Private communication.
Johansson, J., 1989. Emission av N20 frfin'en fluidb/idd i laboratorieskala (N20 emission from a laboratory-scale fluidised bed.) Arbetsrapport-Technical Note, Studsvik
EP-89/46, Sweden (in Swedish).
Wendt, J.O.L. and Ekmann, J.M., 1975. Effect of fuel sulfur species on nitrogen oxide
emissions from premixed flames. Combust. Flame, 25:355 360.
Tseregounis, S.I. and Smith, O.I., 1983. Enhancement of fuel-nitrogen oxidation by
fuel-sulfur in fuel-rich flames. Comb. Sci. Technol., 30:231 239.
Corley, T.L. and Wendt, J.O.L., 1984. Postflame behaviour of nitrogenous species in the
presence of fuel sulphur. II. Rich CH4/He/O2 flames. Combust. Flame, 58:141 152.
Wendt, J.O.L., Wootan, E.C. and Corley, T.L., 1983. Postflame behavior of nitrogenous
species in the presence of fuel sulphur. I. Rich, moist CO/Ar/O2 flames. Combust.
Flame, 49:261 274.
Webster, P. and Walsh, A.D., 1965. The effect of sulphur dioxide on the second pressure
limit of explosion of hydrogen oxygen mixtures. 10th Symp. (Int.) on Combustion, The
Combustion Institute, Philadelhia, PA, p. 463.
Fenimore, C.P. and Jones, G.W., 1965. Sulfur in the burnt gas of hydrogen oxygen
flames. J. Phys. Chem., 69:3593 3597.
Kallend, A.S., 1967. Influence of SO2 on chemiluminescence and atom recombination in
hydrogen flames. Trans. Faraday Soc., 63:2442 2451.
Halstead, C.J. and Jenkins, D.R., 1969. Sulphur-dioxide-catalyzed recombination of
radicals in premixed fuel-rich h y d r o g e n + o x y g e n + n i t r o g e n flames. Trans Faraday
Soc., 65:3013 3022.
Wdjtowicz, M.A., Oude Lohuis, J.A., Tromp, P.J.J. and Moulijn, J.A., 1991. N20
formation in fluidised-bed combustion of coal. In: E.J. Anthony (Ed.), Proc. 11th Int.
Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 1013 1020.
De Soete, G.G., 1989. Internal IFP report No. 36752 (quoted in De Soete, G.G., 1992). Heterogeneous N20 reactions related to combustion. Proc. 5th Int, Workshop on Nitrous
Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. KL-4-1.
Baumann, H., 1992. Private communication.
Gulyurtlu, I., Costa, M.R., Esparteiro, H. and Cabrita, I., 1991. The study of homogeneous and heterogeneous reactions involving N20 and NO~ during fluidized bed combustion of coal particles. Paper presented at the conference on fluidized bed combustion,
December 1991, London, UK.
Gulyurtlu, I., Esparteiro, H. and Cabrita, I., 1992. The study of the reactions of N20 and
NO., with char in a fluidised bed combustor. Proc. 5th Int. Workshop on Nitrous Oxide
Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 5-5-12
Mochizuki, M., Koike, J. and Horio, M., 1992. The mechanisms of N,O formation from
fluidized bec char combustion. Proc. 5th Int. Workshop on Nitrous Oxide Emissions
(NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 5-3-1.
Iisa, K., Salokoski, P. and Hupa, M., 1991. Heterogeneous formation and destruction of
nitrous oxide under fluidized bed combustions. In: E.J. Anthony (Ed.), Proc. 11th Int.
Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 1027 1033.

68
133

134

135

136

137

138
139
140
141

142
143
144
145
146
147

148

149

150

151

152
153

154

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


Miettinen, H., Str6mberg, D. and Lindquist, O., 1991. Tile infuence of some oxide and
sulphate surfaces on N20 decomposition. In: E.J. Anthony (Ed.), Proc. l l t h Int. conf. on
FBC. Am. Soc. Mech. Eng., New York, pp. 999 1003.
Wojtowicz, M.A., Pels, J.R. and Moulijn, J.A., 1991. Solid-catalysed N20 destruction in
combustion of coal. Proc. 1991 Int. Conf. Coal Sci. Butterworth-Heinemann, Oxford,
UK, pp. 452 455.
Naruse, I., Imanari, M., Koizumi, K. and Ohtake, K., 1992. N20 formation and destruction
characteristics in bubbling fluidized bed combustion. Proc. 5th Int. Workshop on Nitrous
Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 8-2-1.
Kimura, N., 1992. N20 emission levels at Wakamatsu 50 MW AFBC plant and N20
control technologies. Proc. 5th Int. Workshop on Nitrous Oxide Emissions
(NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 8-3-1.
Klein, M., K6ser, H. and Rotzoll, G., 1992. Catalytic N20 destruction by alkaline earth
oxides in flue gases and fluidized bed combustors. Proc. 5th Int. Workshop on Nitrous
Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 5-1-1.
Golodets, G.I., 1983. Heterogeneous Catalytic Reactions Involving Molecular Oxygen.
Studies in Surface Science and Catalysis 15. Elsevier, Amsterdam, pp. 200 213.
Riekert, L., Menzel, D. and Staib, M., 1965. Proc. 3rd Int. Congress on catalysis. North
Holland, Amsterdam, p. 387.
Stone, F., 1955. In: W.E. Garner (Ed.), Chemistry of the Solid State. Butterworths,
London.
Saito, Y., Yoneda, Y. and Makishima, S., 1961. Structure-insensitive catalytic activity
as revealed by decomposition of nitrous oxide. Actes du 2e Congress International de
Catalyse. l~ditions Technip, Paris, Vol. 2, pp. 1937- 1954.
Winter, E.R.S., 1970. The decomposition of nitrous oxide on metallic oxides. Part II. J.
Catal., 19:32 40.
Schwab, G.-M. and Staeger, R., 1933. The effect of metallic oxides in decomposing nitric
oxide and their place in the periodic table. Z. Phys. Chem., B21: 65.
Schwab, G.-M. and Staeger, R., 1934. The action of mixed catalysts in the decomposition
of nitrous oxide IV. Z. Phys. Chem., B25: 418.
Gay, I.D., 1970. Catalytic decomposition of nitrous oxide (N20) and oxygen desorption
spectra on nickel oxide. J. Catal., 17: 245.
Winter, E.R.S., 1959. Decomposition of nitrous oxide upon doped nickel oxides. Disc.
Faraday Soc., N 28: 183.
De Soete, G.G. and Nastoll, W., 1992. Catalytic nitrogen chemistry on CaO at fluidized
bed temperature conditions. 10th Members Conf. of IFRF, 13-15 May, 1992, Noordwijkerhout, The Netherlands.
De Soete, G.G., 1992. Unpublished IFP document (quoted in de Soete, G.G., 1992).
Heterogeneous N20 reactions related to combustion. Proc. 5th Int. Workshop on
Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3. 1992, Tsukuba, Japan,
p. KL-4-1.
De Soete, G.G., 1992. Internal IFP report No. 39662 (quoted in de Soete, G.G., 1992).
Heterogeneous N20 reactions related to combustion. Proc. 5th Int. Workshop on Nitrous
Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. KL-4-1.
De Soete, G.G., 1992. Heterogeneous N20 reactions related to combustion. Proc. 5th
Int. Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992,
Tsukuba, Japan, p. KL-4-1.
~mand, L.E. and Andersson, S., 1989. Emissions of nitrous oxide (N20) from fluidized
bed boilers. In: A.M. Manaker (Ed.), Proc. 10th Int. conf. on fluidized bed combustion,
San Francisco, CA. Am. Soc. Mech. Eng., New York, Vol. 1, pp. 49-56.
Aho, M.J. and Rantanen, J.T., 1989. Emissions of nitrogen oxides in pulverized peat
combustion between 730 and 900C. Fuel, 68:586 590.
Aho, M.J., Rantanen, J.T. and Linna, V.L., 1990. Formation and destruction of
N20 in pulverized fuel combustion environments between 750 and 970C. Fuel,
69:957 961.
Gibbs, B.M. and Hampartsoumian, E., 1980. The influence of fuel nitrogen and volatiles
content on the NO emission from the fluidized combustion of solid fuels. Inst. Energy
Symp. Ser. (London), Number: Fluid. Combust.: Syst. Appl., 4:V/2/1-V/2/7.

M.A. Wdjtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


155
156

157

158

159

160

161

162

163

164

165

166

167

168

169

170
171

172
173
174

69

Oude Lohuis, J.A., Tromp, P.J.J. and Moulijn, J.A., 1992. Parametric study of N20
formation in coal combustion. Fuel, 71:9 14.
Gustavsson, L. and Leckner, B., 1991. N20 reduction with gas injection in circulating
fluidized bed boilers. In: E.J. Anthony (Ed.), Proc. 11th Int. Conf. on FBC. Am. Soc.
Mech. Eng., New York, pp. 677 685.
Hiltunen, M., Kilpinen, P., Hupa, M. and Lee, Y.Y., 1991. N20 emissions from CFB
boilers: experimental results and chemical interpretation. In: E.J. Anthony (Ed.), Proc.
l l t h Int. Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 687 694.
Shimizu, T., Tachiyama, Y. and Souma, M., 1991. Emission control of NOx and N_,O of
bubbling fluidized bed combustor. In: E.J. Anthony (Ed.), Proc. l l t h Int. Conf. on FBC.
Am. Soc. Mech. Eng., New York, pp. 695 700.
Bramer, E.A. and Valk, M., 1991. Nitrous oxide and nitric oxide emissions by fluidized
bed combustion. In: E.J. Anthony (Ed.), Proc. l l t h Int. Conf. on FBC., Am. Soc. Mech.
Eng., New York, pp. 701 707.
Braun, A., Bu, C., Renz, U., Drischel, J. and K6ser, H.J.K., 1991. Emission of NO and
N20 from a 4 MW fluidized bed combustor with NO reduction. In: E.J. Anthony (Ed.),
Proc. 11th Int. Conf. on FBC., Am. Soc. Mech. Eng., New York, pp. 709 717.
Brown, R.A. and Muzio. L., N20 emissions from fluidized bed combustion. In:
E.J. Anthony (Ed.), Proc. l l t h Conf. on FBC,, Am. Soc. Mech, Eng., New York,
pp. 719 724.
Jahkola, A., Lu, Y. and Hippinen, I., 1991. The emission and reduction of NOx and N20
in PFB-combustion of peat and coal. In: E.J. Anthony (Ed.), Proc. 11th Int. Conf. on
FBC, Am. Soc. Mech. Eng., New York, pp. 725 730.
Moritomi, H., Suzuki, Y., Kido, N. and Ogisu, Y., 1991. NO., formation mechanism of
circulating fluidized bed combustion. In: E.J. Anthony (Ed.), Proc. 11th Int. Conf. on
FBC., Am. Soc. Mech. Eng., New York, pp. 1005 1011.
Cabrita, I., Costa, M.R., Esparteiro, H. and Gulyurtlu, I., 1991. Determination of overall
fuel-N balance during fluidised bed coal combustion. In: E.J. Anthony (Ed.), Proc. 11th
Int. Conf. on FBC., Am. Soc. Mech. Eng., New York, pp. 985 990.
Wallman, P.H., Ivarsson, E.L. and Carlsson, R.C.J., 1991. NO~ and N20 formation in
pressurized fluidized-bed combustion of coal. In: E.J. Anthony (Ed.), Proc. l l t h Int.
Conf. on FBC., Am. Soc. Mech. Eng., New York, pp. 1021 1025.
Fujiwara, N., Yamamoto, M,, Nishiyama, A. and Kimura, N., 1992. Proc. 5th Int.
Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992,
Tsukuba, Japan, p. 8-4-1.
Andries. J. and Hein, K.R.G., 1992. N20 emissions from pressurized fluidized bed
combustion of coal. Proc. 5th Int. Workshop on Nitrous Oxide Emissions
(NIRE/IFP/EPA/SCEJ), July 1 3, 1992, Tsukuba, Japan, p. 8-5-1.
Lu, Y., Hippinen, h and Jahkola, A., 1992. A comparison of different parameters in
reducing of N20 emissions in pressurized fluidized bed combustion. Proc. 5th Int.
Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992,
Tsukuba, Japan, p. 8-6-1.
Young, B.C., Collings, M.E. and Mann, M.D., 1992. Pilot-seale studies on N20 emissions, coal properties, and conditions in a circulating fluidized-bed combustor. Proe.
5th Int. Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1-3, 1992,
Tsukuba, Japan, p. 9-2-1.
Lu, Y., Jahkola, A., Hippinen, I. and Jalovaara, J., 1992. The emissions and control of
NO., and N20 in pressurized fluidized bed combustion. Fuel, 71:693 699.
Khan, T., Lee, Y.Y. and Brown, R.A., 1992. N,O bench scale studies. Proc. 5th Int.
Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992,
Tsukuba, Japan, p. 9-4-1.
Johnsson, J.E., 1988. Modeling of NO, formation in fluidized bed combustion. Proc. 9th
Int. Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 435 442.
~mand, L-E. and Leckner, B., 1991. Infuence of fuel on emission of nitrogen oxides (NO
and N20) from an 8-MW fluidized bed boiler. Combust. Flame, 84:181 196.
~mand, L.-E. and Leckner, B., 1990. The role of fuel volatiles for the emission of
nitrogen oxides from fluidized bed boilers A comparison between designs. 23rd Symp.
(Int.) on combustion. The Combustion Institute, Philadelphia, PA, pp. 927 933.

70
175

176

177

178

179

180

181
182
183

184
185
186

187
188
189

190

191
192
193

194
195

M.A. W6jtowicz et al./Fuel Processing Technol. 34 (1993) 1 71


Hirama, T., Tomita, M., Horio, M., Chiba, T. and Kobayashi, H., 1984. In: D. Kuni and
R. Toei (Eds.), 4th Int. Conf. on fluidization. Engineering Foundation, New York, Book
No. 83 83069, Vol. 1, p. 467.
~mand, L.-E., Leckner, B. and Dam-Johansen, K., 1992. Influence of SO2 on the
NO/N20 chemistry in fluidized bed combustion. Part I: Full-scale experiments. Fuel,
72:557 564.
Moritomi, H., Suzuki, Y., Kido, N. and Ogisu, Y., 1991. NO~ emission and reduction
from circulating fluidized bed combustor. In: P. Basu, M. Horio and M. Hasatani (Eds.),
Circulating fluidized bed technology III, Proc. 3rd Int. Conf. on circulating fluidized
beds, Oct. 15 18, 1990, Nagoya, Japan. Pergamon Press, Oxford, UK., pp. 399-404.
Shimizu, T., Tachiyama, Y., Kuroda, A. and Inagaki, M., 1992. Effect of SO2 removal by
limestone on NO., and N20 emissions from a bubbling fluidized-bed combustor. Fuel
71:841 844.
Shimizu, T., Tachiyama, Y., Fujita, D., Kumazawa, K., Wakayama, O., Ishizu, K.,
Kobayashi, S., Shikada, S. and Inagaki, M., 1992. Effect of limestone feed on N20
emission from fluidized bed combustors. Proc. 5th Int. Workshop on Nitrous Oxide
Emissions (NIRE/IFP/EPA/SCEJ), July l 3, 1992, Tsukuba, Japan, p. 8-1-1.
Gourichon, L., 1990. N20 measurements in various combustion systems.
LNETI/EPA/IFP European workshop on the emission of nitrous oxide. LNETI, Lisbon,
Portugal, pp. 187 190.
Schaub, G., Reimert, R. and Albrecht, J., 1989. In: A.M. Manaker (Ed.), Proc. 10th Int.
Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 685 691.
Mj6rnell, M., Leckner, B., Karlsson, M. and Lyngfelt, A., 1991. In: E.J. Anthony (Ed.),
Proc. l l t h Int. Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 655 664.
Hirama, T., Takeuchi, H. and Horio, M., 1987. Nitric oxide Emission from Circulating
Fluidized-bed Coal Combustion. In: J.P. Mustonen (Ed.), Proc. 9th Int. Conf. on FBC.
Am. Soc. Mech. Eng., New York, pp. 898-905.
Senary, M.K., 1991. NO~ emission studies in fluidized-bed combustion. In: E.J. Anthony
(Ed.), Proc. 11th Int. Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 1397 1400.
Svoboda, K. and Hartman, M., 1991. Formation of NOx in fluidized bed combustion of
model mixtures of liquid organic compounds containing nitrogen. Fuel, 70:865 871.
De Soete, G.G. and Nastoll, W., 1991. Internal IFP report No. 39362 (quoted in De Soete,
G.G., 1992). Heterogeneous N20 reactions related to combustion. Proc. 5th Int. Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July l 3, 1992, Tsukuba,
Japan, p. KL-4-1.
Cooper, D.A., Ghardashkani, S. and Ljungstr5m, E.B., 1989. Decomposition of ammonia
over calcined and sulfated limestone at 725 950~C. Energy Fuels, 3: 278.
Lee, Y.Y., Sekthira, A. and Wong, C.M., 1985. In: Proc. 8th Int. Conf. on FBC. U.S.
Department of Energy, pp. 1208 1218.
Lee, Y.Y., Soares, S.M.S. and Sekthira, A., 1987. The effects of sulfated limestones on
the ammonia nitric oxide oxygen reaction. In: J.P. Mustonen (Ed.), Proc. 9th Int.
Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 1184 1187.
Dam-Johansen, K., ~mand, L.-E. and Leckner, B., 1992. Influence of SO2 on the
NO/N20 chemistry in fluidized bed combustion. 2. Interpretation of full-scale observations based on laboratory experiments. Fuel, 72:565 571.
Johnsson, J.-E. and Dam-Johansen, K., 1991. In: E.J. Anthony (Ed.), Proc. 11th Int.
Conf. on FBC. Am. Soc. Mech. Eng., New York, pp. 1389 1396.
Chan, L.K., Saroflm, A.F. and Be~r, J.M., 1983. Kinetics of the NO-carbon reaction at
fluidized bed combustor conditions. Combust. Flame, 52:37 45.
Adams, R.C., Aul, E.F., Kulkami, S., McAllister, R.A. and Margerum, S., 1986. Inprocess control of nitrogen and sulfur in entrained-bed gasifiers. EPA/600/7-86/051.
Radian Corporation, Research Triangle Park, NC.
Alexanderson, V. and Sherman; L.M., 1949. Cyanamides. In: Kirk-Othmer (Ed.), Encyclopedia of chemical technology, Vol. 4. Interscience, New York, pp. 663 675.
Lepp/ilahti, J., Simmell, P. and Kurkela, E., 1991. Catalytic conversion of nitrogen
compounds in gasification gas. Fuel Processing Technol. 29:43 56,

M.A. Wojtowicz et al./Fuel Processing Technol. 34 (1993) 1 71

71

196 Groeneveld, M.J., Boxhoorn, G., Kuipers, H.P.C.E., Van Grinsven, P.F.A., Gierman, H.
and Zuiderveld, P.L., 1988. Preparation, characterization and testing of new V/Ti/SiO 2
catalysts for denoxing and evaluation of Shell Catalyst S-995. Paper 223-A, Proc. 9th
Int. Congress on Catalysis, Vol. 4, Calgary, Canada, p. 1743.
197 Zellinger, G. and Tauschitz, J., 1989. Betriebserfahrungen mit der nichtkatalytischen
Stickoxidreduktion in der Dampfkraftwerken den 0sterreichischen Draukraftwerke
AG. In: Proc. of the VGB Knoferenz Kraftwerk und Umwelt 1989, 26 27 April, 1989,
Essen, FRG, pp. 181 186 (in German).
198 Zellinger, G., 1990. Emission data of lignite fired boilers: Influence of different DeNOxprocesses on N20 emissions. LNETI/EPA/IFP European workshop on the emission of
nitrous oxide. LNETI, Lisbon, Portugal, pp. 191 194.
199 Muzio, L.J., 1990. N20 formation and emissions from selective non-catalytic NOx
reduction processes. LNETI/EPA/IFP European workshop on the emission of nitrous
oxide. LNETI, Lisbon, Portugal, pp. 195 201.
200 Hjalmarsson, A.-K., 1992. Interactions in emissions control for coal-fired plants.
IEACR/47, IEA Coal Research, London, UK.
201 Leckner, B., Karlsson, M., Dam-Johansen, K., Weinell, C.-E., Kilpinen, P. and Hupa,
M., 1991. Influence of additives on selective noncatalytic reduction of NO with NH3 in
circulating fluidized bed boilers. Ind. Eng. Chem. Res., 30:2396 2404.
202 Caton, J.A. and Siebers, D.L., 1988. Comparison of nitric oxide removal by cyanuric
acid and by ammonia. SAND-88-8997, Sandia National Laboratories, Albuquerque,
NM.
203 Persson, K., 1990. N20 emissions from coal fired boilers during thermal NOx reduction
tests: Experiences from SWEDCO-supported projects. LNETI/EPA/IFP European
workshop on the emission of nitrous oxide. LNETI, Lisbon, Portugal, pp. 163 170.
204 Hjalmarsson, A.-K. and Soud, H.N., 1990. Systems for controlling NOx from coal
combustion. IEACR/30, IEA Coal Research, London, UK.
205 Jacobs, J. and Hein, K.R.G., 1988. The importance of N20 in relation to nitrogen oxide
emissions. VGB Kraftwerkstechnik, 67(8): 742 744.
206 BWK, 1985. Keine NzO durch Japanische Entstickung-Katalysatoren. BrennstoffWfirme-Kraft, 37(11): 478 (in German).
207 De Soete, G.G., 1990, Nitrous oxide formation and destruction by industrial NO abatement techniques including SCR. Revue de I'Institut Franqais du Petrole, 45(5): 663 682
(in French).
208 Topsoe, N.-Y., Slabiak, T., Clausen, B.S., Srnak, T.Z. and Dumesic, J.A., 1992. Influence
of water on the reactivity of vanadia/titania for catalytic reduction of NOx. J. Catal.,
134:742 746.
209 Singoredjo, L., 1992. Low Temperature Selective Catalytic Reduction (SCR) of Nitric
Oxide with Ammonia. Ph.D. Thesis, University of Amsterdam, Amsterdam, The Netherlands.
210 Gavin, D.G. and Dorrington, M.A., 1991. Factors in the conversion of fuel nitrogen to
nitric and nitrous oxides during fluidised bed combustion. Proc. 1991 Int. Conf. Coal
Sci., Butterworth-Heinemann, Oxford, UK, pp. 347 350.
211 Braun, A., 1990. Emission of NO and N20 from a 4 MW fluidized bed combustor. 21st
IEA-AFBC Meeting, Belgrade, Serbia, Nov. 7 10, 1990.
212 Suzuki, Y., Moritomi, H., Kido, N., Ikeda, M., Suzuki, K. and Torikai, K., 1992. N20
formation from char and heterogeneous reactions with char and CaO. Proc. 5th Int.
Workshop on Nitrous Oxide Emissions (NIRE/IFP/EPA/SCEJ), July 1 3, 1992,
Tsukuba, Japan, p. 5-4-1.
213 Suzuki, Y., Moritomi, H. and Kido, N., 1991. On the formation mechanism of N20
during circulating fluidized bed combustion. Presented at the 4th SCEJ Symp. on
Circulating Fluidized Beds. Soc. Chem. Eng. Jpn., Tokyo, Japan.
214 Leckner, B. and Gustavsson, L., 1991. Reduction of N20 by gas injection in circulating
fluidized bed boilers. J. Inst. Energy, 64: 176-q82.

También podría gustarte