Sie sind auf Seite 1von 23

Pro9. Enerfy Combust.Sci.. 1976, Vol.

1, pp. 111 133.

Pergamon Press.

Printed in Great Britain

GAS TURBINE ENGINE POLLUTION


A. M. MELLOR*
The Combustion Laboratory, School of Mechanical Enoineering, Purdue University, West Laj?Lyette,
Indiana, U.S.A.

Abstract--Recently set U.S. Environmental Protection Agency standards and Air Force goals for aircraft
jet engine emissions of unburned hydrocarbons (HC), CO and NOx are reviewed in terms of the contribution
of air transport to the overall pollution problem. Suggested design modifications for gas turbine combustors
to meet these requirements are discussed via a generalised physical model of the liquid fuel spray
combustion process which is consistent with both engine and combustor exhaust plane pollution measurements and data obtained within burners. This model is used to show why some analytical/numerical
combustor modeling efforts succeed in correlating emissions and may eventually become useful design
tools for next-generation, low emissions Brayton cycle engines.

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

Smoke in the exhaust plumes of the then new commercial turbojet engines resulted in a survey of aircraft
emissions by the Los Angeles County Air Pollution
Control District, the results of which were published by
R. E. George and R. M. Burlin in 1960. 31'v4 Later
work 5v'~4 concluded that at most aircraft contribute
only 1 or 2% of the total pollution from all sources in
Los Angeles County, and more current estimates s2
verify these figures nationwide in the U.S.A. (Table 1).
Why then the concern over aircraft emissions, culminating in the U.S. Environmental Protection Agency's
issuance in 1973 of standards for civil piston- and
turbine-powered aircraft 27 and the U.S. Air Force's
guidelines for military jets ? 12,13.41
TABLE 1. Aircraft Contributions to Air Pollution, % of
Tota174,60.9.s2

CO
United States, 1968
2.4
New York City, 1968 0.6
Los Angeles, 1970
1.5
San Francisco, 1970 1.0
San Francisco, 1985 10.9

HC

NOx Particulate SOx

0.9
0.6
3.2
1.5
3.5

1.7
0.2
1.4
1.3
11.6

0.1
0.2
11.5
6.6
9.3

0.03
-2.0
1.0
2.2

In urban areas with large airports, aircraft emissions


now represent a much more sizable contribution to the
regional picture, particularly when considered in terms
of the increases in air travel traffic since 1960 and the
standards which have been applied to other sources
such as automobiles and power plants. For example,
by 1975 in the metropolitan Los Angeles area, aircraft,
if uncontrolled, are projected 26'47 to account for 6.4%
of the carbon monoxide, 2.8~o of the unburned hydrocarbons (HC) and 2.8% of the oxides of nitrogen (NOx)
released into the atmosphere from all sources; by 1980,
these percentages become 13.6, 2.5 and 5.7, respectively.
Similar figures are shown for San Francisco in 1985 in
Table 1. Also, within and near such airports pollution
from aircraft is estimated to be largely responsible for

1. E P A Aircraft Emissions Standards


The basic unit of the EPA standards is the landing
and takeoff (LTO) cycle, which as can be seen from
Table 2 simulates civil ground operations as well as
those occurring in the immediate area of the airport.
TABLE2. LTO Cycle for the Jumbojet Engine 2v't3

* Professor of Mechanical Engineering.


t This section is based on a note which appeared in the
1974 Fall issue of California Air Environment.
JPECS Vol. 1, Nos. 2/3

the surpassing of the U.S. national ambient air quality


standards for C O at least 39 times during the winter of
1970 at Los Angeles International Airport/6
To date then, most concern has been voiced about
aircraft at and near airports, and in response EPA has
defined their regulations in terms of aircraft operations
on the ground and below about 0.9 km altitude. 26'2~'47
A related but separate issue is the high-altitude stratospheric ( > 10 km) emissions of the supersonic transport,
but characterization of SST emissions and their
possible impact (and appropriate instrumentation for
their measurement) is still an active area of research. 1
Only recently has EPA proposed standards, 28 and these
are also oriented toward landing and takeoff procedures.
Although piston-powered aircraft do contribute
substantially to the civil aircraft emissions in Los
Angeles County itemized above, design modifications
and control techniques for this type of engine are
discussed elsewhere; 43 thus here we concentrate on
combustor technology for and pollution from gas
turbine engines. The potential use of such prime movers
in heavy-duty vehicles, and possibly automobiles, and
very encouraging thermal efficiency projections sl for
electric power generating stations operating on a
combined Brayton-Rankine cycle point to more widespread application of these continuous combustors in
the future.

111

Engine setting

Power setting,
}0 of maximum rated thrust

Time in
mode, min

Taxi/Idle (Out)
Takeoff
Climbout
Approach
Taxi/Idle (In)

a
100
85
30
a

19.0
0.7
2.2
4.0
7.0

a AS specified by engine manufacturer.

112

A.M. MELLOR

The cycle is defined to reflect different engine sizes;


that for the Pratt & Whitney JT9D jumbojet engine
is given in Table 2 and is similar to those for other
engine classes. To be controlled during this cycle are
emissions of unburned fuel and other hydrocarbons
(HC), CO, NO~ and smoke. Sulphur oxides are not
considered, since aviation fuels contain only minimal
amounts of sulphur, a specification to minimize corrosion of the combustor and turbine. All standards will
take effect in stages but by 1981 (in recognition that the
lifetimes of existing engines and development times of
advanced engines are long); "retrofit" standards for
existing engines are limited to date to smoke emissions
and fuel venting. 47

where ~Ib is the combustion efficiency, EI X is the


emissions index of species X (g X/kg fuel) and Qx is the
lower heating value of species X. If QHC is taken equal
to Qfue~ for JP4 fuel and numerical values are
substituted, then
qb = 1 -- [0.232 (El CO) + (El CxHy)] x 10-

where the emission indices at engine idle are used for


comparison with the idle combustion efficiency goals.
As noted in section 1.1 above, most HC and CO are
emitted at engine idle, so that specification of these
emissions only at idle is satisfactory.
For thejumbojet engine, even though NOx emissions
are distributed over the takeoff, climbout and approach

TABLE 3. Comparison of 1979 EPA Standards and USAF Goals 13

Idle combustion efficiency, ~o


EPA engine
class
T1

T2, T3
P2

Manufacturer
and engine

EPA
standard S

General Electric
J85
AiResearch
TFE-731
Pratt & Whitney
JT3D
JT9D
AiResearch
TPE-331
Allison T56

NO~ emissions index at takeoff

USAF
goat

EPA
standard a

USAF
goal

99.6

98.0

7.0

6.0

98.88

99.0

l 0.0

10.0

99.2
99.0

98.0
99.0

7.0
10.5

31.5/10.5 b

99.2
98.4

99.0
99.0

7.0
20.5

7.5
10.0

10.5/3.5 b

a EPA parameter in lb/10001b thrust hr/LTO cycle for CO and HC converted to equivalent combustion
efficiency or NO~EI.
b If water injection can be applied, a 75~ reduction ig anticipated.
Note that the L T O cycle involves both extremes of
engine operation, low-power idling and high-power
takeoff (Table 2). The former is primarily responsible
for emissions of HC and CO during the cycle (97.3~o
and 94.3~o, respectively, for jumbojet engines14,13),
while NOx and smoke concentrations are more concentrated at the takeoff and climbout conditions.14'13

2. USAF Aircraft Emission Goals


Because military aircraft design philosophy is
different from that of civil airplanes, and because
military ground operations are less reproducible than
their civil counterparts, the U.S. Air Force has chosen
to define their goals in terms of reductions from present
uncontrolled
emissions.
Also, Blazowski and
Henderson 13 note that EPA's use of emissions units as
mass pollutant per unit force thrust per LTO cycle
introduces specific fuel consumption, and thus engine
design, into their parameter (see also section IliA).
Tradeoffs between performance and pollution are not
acceptable for military aircraft, and thus Air Force
goals are defined only in terms of absolute emission
levels. The HC and C O goals are combined in
calculating combustion effciency:
qb--- 1

(El CO)Qco + (El CxHy)QHc


Qf~el 10 3

modes in the LTO cycle (26.7, 53.6 and 10.5~ of total,


respectively14'~3), for simplicity and convenience only
the emissions at takeoff power are involved in the goals.

3. Comparison of Standards and Goals


Blazowski and Henderson ~3 have compared the
standards and goals insofar as possible; their results
shown in Table 3 demonstrate good agreement for HC
and CO emissions, but the NOx goals of the Air Force
are less stringent for the T2 class if water injection is
not used. Smoke goals are defined in both cases to make
aircraft plumes invisible)3

4. Origin of Pollutants within Gas Turbine Combustors


Both the Environmental Protection Agency and the
Air Force have indicated concern over emissions of
HC, CO, NOx and smoke: all of these pollutants are
combustion generated, in aircraft combustors which to
date use extremely similar designs (Fig. 1). Whether a
single combustor can, or annular or can-annular
geometries are used,* as shown in Fig. 1, the cross* In a small engine a single can combustor and fuel injector
may suffice; for larger engines either a number of injectors
and cans, arranged in a circle corresponding approximately
to the turbine diameter (can-annular), or a single annular
combustor with several fuel injectors is used.

Gas turbine eng!ne pollution

FUllLE
;
I PRIMARY COMBUSTIONI
~,
ZONE
I

VAPORIZATION
ZONE

SECONDARY tp
COMBUSTfON
ZONE

=
~

SECONDARYI
:
COMBUSTION I
ZONE
I DtLUTION ZONE I

WALL
CENTRAL
IMPINGEMENT
RECIRCULATION RECIRCULATION RECIRCULATION
ZONE
ZONE
ZONE

discussed in section V. The more illuminating contour


maps and flow models obtained by probing directly
within combustors are shown generally consistent with
Fig. 1 (section VI) and then compared with available
numerical/calculational models of the gas turbine
combustion process (section VII). First let us review
how conventional turbine combustor configurations
have developed and how the gaseous pollutants HC,
C O and N O are formed in such burners.

FIG. I. Typical combustor cross-section showing the


various flow r~gimes.
section of the combustor is generally considered to
consist of three major zones (primary, secondary, and
dilution), although their physical boundaries are rarely
well defined. The particular philosophy which led to the
evolution of this conventional-type combustor will be
indicated in section II, and it will be shown that the
particular facet of the design important for pollution
characterization is the volume adjacent to the
individual fuel injector in the primary zone; here the
liquid fuel flame is established and poor mixing with
air leads to smoke,* while the high temperature at
which combustion occurs (the average equivalence
ratio is near stoichiometric in conventional primary
zones) causes nitric oxide (NO)t formation. Since
these temperatures are too high for the turbomachinery
which follows in the engine, excess air is added via
penetration jets and/or film-cooling slots (Fig. 1) to cool
the working fluid. If this is done improperly, then HC
and CO reactions are quenched.
EPA has proposed minor changes in combustor
design and engine operation for existing gas turbines
which should assist in meeting their emission standards
and which, as applied to conventional burners, are the
subject of section III. However, because more
sophisticated modifications will probably be necessary,
in section IV advanced design concepts are evaluated
using reported emissions levels wherever possible.
Some of these consist of pollutant concentration
measurements at the burner or engine exhaust plane,
and the resulting correlations or physical models are
* Since smoke control techniques have been developed
empirically and are well known (see, for example,
Lefebvre55), they will not be discussed in the following.
t As noted in previous sections, the EPA standards and
USAF goals are written in terms of oxides of nitrogen, NOx
(nitric oxide, NO, plus nitrogen dioxide, NOz). Most of the
NOx is emitted by aircraft as NO, but sampling and measurement uncertainties have made quantitative determination of
the NO2 concentration difficult. 94 It appears to be most
significant when high compressor pressure ratio engines are
operated at low power settings. Cernansky and Sawyer is
and Merryman and Levy67 have recently suggested that the
reaction N O + H O z ~ NO2 + O H is responsible for NO/
formation in flames, and the high combustion pressures and
moderate temperatures in such engines at idle would favor
the HOz radical.
Since photochemical smog induction time depends on the
ratio of NO to NO2 in the atmosphere, 1 and since idle
engine operation is an important portion of the LTO cycle
(Table 2), both this ratio and total NOx are significant in
determining the aircraft contribution to air pollution.

113

lI. C O N V E N T I O N A L

COMBUSTOR

DESIGN

Current conventional combustor design is based on


the extremely important requirements of stable and
efficient combustion (the combustion efficiencies at idle
listed in Table 3 are the lowest encountered over the
engine operating range) and easy ignition at any
altitude the airplane may encounter.

1. Evolution of the Conventional Combustor


Lefebvre 53 has shown the design genesis and
rationale particularly well, and his exposition is shown
in Fig. 2. Figure 2A shows the most simple combustor
imaginable, where liquid fuel is simply sprayed into a
duct of constant cross-sectional area. Lefebvre 53 notes
that this geometry is impractical, however, because of
the high air velocities coming directly from the air
compressor (which make flame stabilization impossible) and the large pressure loss due to combustion

~
(A)

FU EL

AIR ~
|

(B)

AIR

EL

~FUEL

(C)

AIR - -

FUEL

(D)

AIR

LOCAL ~=l

75%

/-

FIG. 2. Stages in the evolution of the conventional aircraft combustion chamber


(from Lefebvre53).

114

A.M. MELLOR

in this configuration. In Fig. 2B a diffuser has been


added to reduce the combustor inlet air velocity to
about 30m/sec, but this is still too large to allow
stabilization of a flame on the fuel injector.
A baffle or bluff-body stabilization scheme is
suggested in Fig. 2C: the flow reversal or recirculation
zone into which the fuel is sprayed allows stabilization
over a wide range of equivalence ratios, but this range is
too narrow to include temperatures sufficiently low for
the downstream turbine.* Thus the total compressor
discharge airflow is split by the use of a liner or flametube, so that in the primary zone the average
equivalence ratio is within the fuel/air flammability
limits and continuous combustion is possible. The flow
within the primary zone still uses a recirculation zone
for stabilization, as shown in Figs. 1 and 2D: reversal
is accomplished by swirling the air injected through the
dome surrounding the fuel nozzle, by sudden area
increases for this airflow, and/or by impingement of
air from penetration jets (see Fig. 1). In this basic
configuration combustion e fficiencies in excess of 99%
can be achieved routinely at engine design operation
(i.e., high air and fuel flow-rates), but at the lower fuel
and air flows characteristic of engine idle, efficiency
drops off (see section 1.2). The fixed air addition
geometry provides sufficient flow reversal and thus
flame stability over the entire engine load range at an
acceptable pressure drop (less than 539.
2. The Conventional Primary Zone and Pollutant
Formation
Typical primary zone flow patterns for the bluff body
and flame-tube combustors are shown in Figs. 3A and
3B. Although the latter may involve air swirl, the major
elements of the two which determine performance
(efficiency and stability) and emissions in conventional
combustors are very similar. The arrows represent the
time-averaged flow patterns, but since the flow is
turbulent, mixing is in reality accomplished via a shear
layer shown at an instant of time schematically in
Fig. 3C for the disc stabilizer.t In this layer fresh air
from the compressor is mixed with the fuel and burned
gas of the centerline recirculation zone. Depending on
the magnitude of the liquid fuel droplet lifetimc (Zeh)
the fuel can be either liquid or vapor when it reaches
the outer edge of the recirculation zone. Definition of
similar characteristic times for the other governing
physical and chemical processes is useful in understanding the complex two-phase turbulent flow with
chemical reaction (see, for example, refs. 33, 64, 99 and
3 as applied to gas turbine combustion).
Note that there are two mixing processes of
importance in this schematic of the primary zone of a
* Overall combustor equivalence ratios are on the order of
0.2: even at the elevated compressor outlet temperatures
and pressures currently in use such mixtures are too dilute
to burn.
t" This flow is axisymmetric and will serve to demonstrate
the essential features of the more complicated threedimensional combustor flow.

(A)

F
U
E
L
//////~///_//////////~
-

(B)

F
U
E
L

~////////////////////i

(C)

A
R
I

~ ~
V_---

FIG. 3. Simplified schematics of various


primary zones.
conventional combustor (Fig. 3C): one involves the
mixing of the fuel vapor eddies with the hot,
recirculating burned gases, and the other that of the
resulting fuel-burned gas mixture with the fresh air in
the shear layer. Characteristic times for these two
mixing processes can be thought of as turbulent eddy
dissipation times or lifetimes; let us call that associated
with the fuel injector D~ and that with the shear layer

Tsl.
Finally, there are two characteristic times associated
with the homogeneous chemistry of the fuel/air combustion. The one pertinent to performance and lowpower emissions of HC and CO is the fuel ignition
delay and burning time ~nc, but this is generally short
compared to the time required for N O formation Tno,99
All of these characteristic times should be thought of
as order of magnitude estimates, since each takes on a
range of values in an actual combustor: the drop
lifetime is a function of initial drop size (typical turbine
injectors do not produce monodisperse sprays), and all
five times depend on local gas temperatures, velocities,
and compositions, which vary considerably throughout
the primary zone. With these limitations in mind one
can accept these times in the proper spirit, as they do
identify the key processes affecting both emissions and
performance. For convenience they are summarized in
Table 4.
An example of their utility is for the bluff-body flame
stabilizer used in turbine afterburners and ramjets.
Liquid fuel can be sprayed into the flow upstream of
the stabilizer so that it has at least partially vaporized
and mixed with the combustion air. In attempting to
explain and correlate data on the variation of approach

Gas turbine engine pollution

I 15

TABLE 4. Characteristic Times for Combustion and Pollutant Formation in Two-phase Turbulent flow
Time

Symbol

Physical or chemical l~rocess

Fuel droplet lifetime


Eddy dissipation time for injected fluid

r~b
rf~

Eddy dissipation time in the shear layer

Zst

Fuel ignition delay and burning time

zh~

NO formation time

r,o

Droplet evaporation and/or combustion


Small-scale turbulent mixing near the fuel
injector in the recirculation zone
Large-scale turbulent mixing between
fresh air and the recirculating burned
gas-fuel mixture
Homogeneous combustion of the fuel to
CO2
Homogeneous kinetics for NO formation

velocity at which the flame is extinguished (the blowoff


velocity) with equivalence ratio, inlet temperature and
pressure, etc., two main theories for stability evolved.
Longwell and co-workers 59 believe that if the m e a n
gas residence time in the recirculation zone exceeds the
requisite fuel burning time there, then a stable flame
will spread into the freestream. Zukowski and
M a r b l e ~3 prefer to associate stability with the shear
layer shown in Fig. 3C; stable combustion is achieved
if the fluid residence time is greater t h a n the burning
time in the shear layer. The residence time should

flames due to either slow evaporation or diffusion


flame combustion of droplets of large diameter.*
Improved mixing (decreased ~ii or ~s~) similarly can
speed the combustion of fuel v a p o r eddies, either with
the hot recirculating b u r n e d gas and air in the
recirculation zone or with fresh air in the surrounding
shear layer; b o t h improved combustion efficiency and
lower N O result. In certain types of advanced combustors to be discussed in section IV ~hc and r.o are
design variables, obtained by premixing the fuel and
air before ignition (so that the equivalence ratio at

TABLE 5. Pollution Reduction by Variations in the Characteristic Times


Characteristic time
and variation

Varied by
for example

Decrease feb

Improving fuel
atomization

Decrease zj.~

High velocity
air injected with fuel

Decrease ~sl

Increase air
flow rate through
primary zone
Increase local
reaction temperature
Decrease local
reaction temperature

Decrease ~h,.
Increase r,o

perhaps be replaced with the eddy lifetime in the shear


layer, since flows of practical interest are turbulent, and
then the flame will be blown off when the a p p r o a c h
velocity is just increased past the point where
"(sl ~

~hc,sl

where Zhc.s~is the fuel ignition delay and burning time


under the conditions in the shear layer. The Z u k o w s k ~
M a r b l e theory can explain the blowoff-relight
hysteresis observed in gas turbine combustors 45 and
recently was shown applicable to lean flammability
limits of a simplified prevaporizing/premixing gas
turbine combustor. 3
The influence of each of the characteristic times
defined in Table 4 upon pollutant emissions is indicated
in Table 5; methods of varying these times are
discussed in section III. Decreases in feb reduce all
three gaseous pollutants by avoiding elongated spray

Effects of the variation on exhaust emissions


(unless otherwise stated)
Reduce HC and CO by providing fuel vapor at a faster rate
and reduce NO by decreasing overall flame length; or
reduce HC, CO and NO by droplet combustion of
decreased duration
With air present near the injector, decrease local HC, CO
NO as a result of increased burning rate through improved
mixing
Reduce HC, CO and NO through more rapid mixing and
combustion in the large-scale shear layer: but HC and CO
can increase as a result of quenching
Decrease HC and CO by increasing their chemical kinetic
rates of consumption
Decrease NO by decreasing its chemical kinetic rate of
production

which combustion occurs in the primary zone is a truly


independent parameter). Then H C and C O emissions
can be decreased by reductions in zhc and N O decreased
by increases in %0.
F o r conditions typical of stable combustion at engine
design loading, Vranos 99 has estimated four of these
times; his results are shown in Tables 6 and 7. zs~ is
estimated from an equation provided by Corrsin 23
"rsl = c ( m L 2 / p ) 1/3

where e is a proportionality constant evaluated from


stirred reactor experiments, m is the fluid mass in the
* Note that Zebis evaluated for either droplet evaporation
droplet combustion. Diffusion or wake flame burning of
individual droplets depends on drop diameter and its
velocity relative to the ambient gas, 87 and the m a y n i t u d e of
z~b does not reflect which is responsible for mass loss from
the drop.
or

l 16

A.M. MELLOR

TABLE 6. Estimated Characteristic Times for Primary Zone


Conditions 99
Average primary zone
rsl, msec
equivalence ratio, ~b,~
0.7
0.8
0.9
1.0
1.1
1.2
1.4

10
l0
l0
l0
l0
10
10

The , msec

0.20
0.17
0.15
0.13
0.16
0.20
0.40

z,,o, msec
330
53
15
7
23
109
> 1000

Varied by changing fuel flow at constant air flowrate.


primary zone, L is a characteristic c o m b u s t o r dimension and P is the power input calculated from the
liner pressure drop and air flowrate. F o r fixed geometry
and flowrate, %~ is constant a n d estimated as 10msec,
where a c o m b u s t o r inlet pressure of 14.67atm and
inlet temperature of 700 K are assumed.

TABLE 7. Estimated Characteristic Times for Liquid Fuel


Injection 99
Mean drop diameter,
[1m

Teb, msec

150
100
50
25
10

9.8
3.6
1.2
0.4
0.2

Zsl, m s e c
10
10
10
10
10

~hc, a msec

0.13
0.13
0.13
0.13
0.13

Taken for stoichiometric combustion.


The burning time 3he is obtained from laminar f a m e
speed and deflagration thickness, and 3,o from the
initial kinetic rate of N O formation compared to N O
concentration at chemical equilibrium at the equivalence ratio in question. D r o p lifetimes are computed
from the "d 2 law" corrected for forced convection and
aerodynamic drag, taking properties of JP5 fuel.
O n the basis of these data Vranos 99 concludes that
the spray combustion process at engine design in a
conventional c o m b u s t o r should resemble a turbulent
diffusion flame since zst >> 3hc; that only when 3no < %t
will N O form to appreciable a m o u n t s (i.e., for
0.95 < 47 < 1.0); and that 3e~ ~< %t so that droplets can
at times be i m p o r t a n t to the combustion process (see
also ref. 63). However, since for 3st he took L typical of
the c o m b u s t o r diameter rather t h a n of an air penetration hole or swirler slot, 9 Vranos's estimates for the
shear layer mixing time could be an order of magnitude
too high. We shall return to his conclusions in sections
V and VI below; but note that they imply that in
conventional combustors r~t and perhaps r~b determine
the emissions of HC, C O and NO, and not the
characteristic h o m o g e n e o u s kinetic times rh,. and
%o. With this in mind let us consider possible emissions
control techniques in terms of these characteristic
times and the simplified primary zone model shown in
Fig. 3C.

Ill. EMISSIONS C O N T R O L T E C H N I Q U E S FOR


C O N V E N T I O N A L BURNERS

E P A 26 has classified control techniques into major


and m i n o r state-of-the-art c o m b u s t o r design modifications. The latter are thought to be applicable to conventional combustors currently in use in civil and
military aircraft, require c o m b u s t o r a n d / o r injector
retrofit and are the topic of this section. As was
suggested in section II, b o t h the liquid phase of the
injected fuel and its initial segregation from the air
(with which mixing must occur before combustion)
force chemical reaction to occur at the highest possible
temperature on the average in a conventional combustor. This in turn generates m a x i m u m values of NO.
Fast and t h o r o u g h fuel/air mixing tends to minimize
smoke, HC, and CO, but too rapid mixing can quench
reactions of the latter two species.* W h e n fuel flow is
increased from the idle to full power setting, conventional pressure-atomizing injectors tend to improve
in atomization; 54 thus to reduce emissions over the
L T O cycle b o t h the heterogeneous effects important
perhaps at idle and the mixing effects controlling at
full power must be altered. The m i n o r modification
control techniques listed in Table 8 ur/fortunately
TABLE 8. Emissions Control Techniques Requiring Minor
Combustor/Injector Modifications 26
EPA designation a

Modification

t 1 -- minor

combustion Minor modification of combustion


chamber redesign
chamber and fuel nozzle to
achieve best state-of-art
emission performance
t4 divided fuel
Provide independent fuel supplies
supply System
to subsets of fuel nozzles to
allow shutdown of one or more
subsets during low-power
operation
t5 water injection
Install water injection system for
short duration use during
maximum power (takeoff and
climb-out) operation
t6 -modifycompressor Increase air bleed rate from
air bleed rate
compressor at low-power
operation to increase
combustor fuel air ratio
t3 is omitted here because it involves fuel drainage
control and is thus not connected with the combustion
process.
address in general only one characteristic time at one
engine power setting and thus probably will prove
unsuitable for meeting EPA's aircraft emissions
standards 9s (see section III.4 below). They may suffice,
however, for meeting Air Force goals.
1. Variation in Fuel Droplet Lifetime and
Small-scale M i x i n 9 Rate (zeb and z fl)

Excessively long burning or evaporation times for the


droplets injected into the liquid fuel spray flame will
* For combustors with large surface area to volume ratios,
quenching by film cooling air injected along the liner walls
can also be important.

Gas turbine engine pollution


lead to incomplete combustion and thus increased
emissions of HC and CO. As noted above, pressureatomizing nozzles function less efficiently in terms of
droplet breakup as the fuel flow is reduced toward the
engine idle condition. One method of decreasing ze~
and/or rli, which involves modification to the fuel
supply system, is control method t4 (Table 8), where
more fine atomization is obtained at idle by shutting off
flow to some nozzles and increasing flow to the remaining nozzles proportionally. The same total fuel flow is
maintained, but each operating injector enjoys a greater
fuel flowrate and thus provides better atomization.
Jones and Grobman 48 show the idle combustion
efficiency gains which can result, but it is not clear that
decreased fuel drop diameters are alone responsible,
since HC and CO emissions are strongly influenced by
quenching.
Control method tl involvinginjectors is also directed
toward decreasing the mean drop diameter and therefore lifetime, for example, by replacing the pressureatomizing nozzles (in predominant use on current
American-made engines) with either airblast or airassist injectors. The pressure atomizers rely on a large
fuel differential injection pressure drop across a single
or double orifice to shatter the liquid fuel into a fine
spray (simplex and duplex nozzles, respectively),
whereas the airblast nozzles use the momentum of the
compressor discharge air, some of which is directed
through the fuel injector, to aid in droplet breakup;
air-assist implies the use of air pressurized to even
higher levels by an auxiliary pump. With a pressureatomizing pilot injector, airblast and air-assist nozzles
can circumvent any flame stabilization difficulties. ~3
Because these injection methods introduce air,
sometimes with swirl, intimately with the liquid
fuel in the vicinity of the injector, the mixing rate
there is increased (rli is decreased) which also tends
to decrease emissions of HC and CO. Many
investigators35'v3'77'24'48'7 have reported this advantage with various types ofairblast and air-assist nozzles.
The effect on NO emissions is not as readily apparent:
increased turbulent intensity and velocities near the
injector may preclude droplet burning and thus
decrease NO, while the presence of air in this region
can lead to locally high NO concentrations. For
example, Pompei and Heywood 8 report that NO
exhaust emissions are independent of atomizing air
differential injection pressure for mean primary zone
equivalence ratios near stoichiometric but decrease
with increases in this parameter for leaner flames.
Other investigatorsz4'48'7 note lower NO emissions
for these nozzles than from pressure atomizers.
There is some evidence, based on concentration
measurements within conventional combustors (see
section VI), that most of the NO emitted in the engine
exhaust forms in the shear layer surrounding the flameholding recirculation zone (see Fig. 3C). For can-type
combustors predominantly of either the film cooled or
the penetration jet type using pressure atomizing
injectors, NO concentration in the immediate vicinity
of the fuel spray, although reaching locally high values,

117

remains to be diluted by most of the engine air flow


and thus contributes only a small percentage of the
total emissions.93'ss As a result, control methods
directed at modifying ~st (rather than %b or zli) should
have a larger effect on NO emissions from conventional
burners.

2. Variation in Large-scale Mixing Rate (%~)


Rearrangement in penetration jet placement and
percent of total airflow passing through the primary
zone can significantly affect the mixing rate in the shear
layer. Decreases in r~t will improve combustion
efficiency and lower NO, since faster mixing will allow
more rapid combustion of all of the fuel (recall
%~ >>rhc) and simultaneously allow less time for NO
formation (recall %z ~<%o).Such variations in flametube
configuration constitute another aspect of EPA's
control method tl. However, related changes possible
in other parameters such as the overall length of the
spray flame and its rate of fresh air entrainment may
obscure these desirable results.
Automative gas turbine combustors require much
simpler design than the typical aircraft combustor
shown in Fig. 1, due to reduced cost requirements,
and thus serve well to exemplify the effects of flametube
design changes in NO. In particular, the GT309
combustor in its original design uses air addition
through louvered holes in the dome in its upstream end,
a film cooling slot surrounding the dome, one row of
small diameter holes in the liner walls for injecting air
into the primary zone and another row of large holes
downstream for dilution air addition? As shown in
Fig. 4, by changing the total airflow to the primary
zone (curves labeled rich and lean primary zone and
early quench), NO emissions were lowered over the
entire engine design range, with as much as a 40~o
reduction at full power; this was accomplished merely
140

120

I00

~0
----0~
__ - - .

o_
I

z
o

s0

- ..43 /

STANDARD BURNER
RICH PRIMARY ZONE
LEAN PRIMARY ZONE
EARLY QUENCH, MODIF.
NO. 2
~' LEAN PRIMARY ZONE AND
EARLY OUENCH, MODIE
NO. I

4o
20

...43
~

,~O
[3
<>

50

60

70

8o

so

~oo

GASIFIER TURBINE SPEED-PERCENT OF DESIGN

FIG. 4. Exhaust NO emissions of modified GT309


burners versus percent of design gasifier turbine speed
(from Wade and Cornelius1).

118

A.M. MELLOR

by changing the location and/or diameter of the holes


referred to above, but any corresponding increase in
HC and CO due to quenching was not reported (see
Wade and Cornelius').
As noted previously, for any type of engine poor
combustion efficiency is generally encountered only at
idle. Therefore maintaining the airflow pattern and
resulting mixing rate occurring at engine design can
increase emissions of HC and CO through quenching
of the combustion process at low power settings.
Control method t6 (Table 8) essentially recognizes the
possibility of a small z~t (advantageous at full power)
quenching HC and CO at idle. By increasing the airflow
bled from the compressor (and thereby decreasing the
airflow to the combustor) this technique increases z~l,
but t h e gain in average combustor residence time
overpowers the poorer mixing, and idle combustion
efficiency can increase (see also section III.l above).
Note that we prefer to interpret changes in mean
primary zone equivalence ratio in conventional burners
(see Table 8, control method t6, and Fig. 4) in terms of
changes in shear layer mixing rate or turbulent eddy
dissipation time in the shear layer around the
recirculation zone (r~). It has been suggested above
that the usual spray combustion process resembles a
turbulent diffusion flame (see also section VI): for such
a flame the burning rate is equal to the mixing rate;
the equivalence ratio at which combustion actually
occurs is on the mean stoichiometric: and thus the
concept of a primary zone equivalence ratio can only
reflect differences in combustion volume or flame
length.
3. Variations in the H o m o g e n e o u s Kinetic" R a t e s

(rhc and r,o)


However, one can also consider how such design
changes will affect the kinetic rates of fuel and CO
oxidation (or Zhc) and NO formation (or z,,,,). In this
view mean primary zone equivalence ratios near
stoichiometric increase all of these rates via their
dependence on temperature. The ability to decrease
NO emissions by water injection into the air flow
upstream of the combustor (control method t5) is best
thought of as increasing r,o (decreasing the NO formation rate) by dilution causing the lowering of peak
temperatures in the primary zone, although the resulting increased mass flow into the primary zone also
decreases z,i (and thus NO formation).
The two characteristic times "Chcand zno do assume
control of the combustion process in the advanced
prevaporizing/premixing burners to be discussed in
section IV, and we shall return to them there.
4. M i n o r Modification to Conventional C o m b u s t o r s and
the Emission Standards

Verkamp et al. 98 consider the emission standards to


consist primarily of a trade-off between CO and NO,
since if the CO standard is met HC emissions are
usually also sufficiently low. To demonstrate this idea

I00

I0

197(3
.REGI
35 :
.50
1.00
I

I0

IOO

NOx El, g NOx/kg fuel

FIG. 5. Emissionstrade-offfor conventional combustors over the


LTO cycle and class T2 1979 EPA
standards as a function of engine
specificfuelconsumption (SFC in lb
fuel/lb thrust hr) (from Verkamp
et al.98).
they present exhaust emissions index of CO versus
NOx emissions index (Fig. 5); the figure also shows
how the EPA standards depend on engine specific fuel
consumption (see section 1.2). The band of idle to
maximum power emissions is obtained from exhaust
measurements for a wide range of Detroit Diesel
Allison, Pratt & Whitney and General Electric engines
(includinglow and high compression ratio, regenerative
and nonregenerative engines, and can, can-annular and
annular combustors). 98 Note that only those unmodified engines having the lowest specific fuel consumption can expect to satisfy the class T2 1979
standards over any portion of their operating range.
These investigators report experimental results of
Allison studies on the potential of the minor modification control techniques discussed above to shift the
band of exhaust emissions shown in Fig. 5 toward the
origin. Control method tl (minor combustor and/or
injector redesigns, including airblast and air-assist
atomization) could accomplish at best a 50% reduction
in NOxEI and CO EI simultaneously; most individual
design changes by themselves yield exhaust emissions
still within the conventionalcombustor emissions band.
Water injection (t5) gave significant decreases in
NOxEI, but is thought impractical due to logistics and
aircraft safety requirements. Verkamp et al. 98 conclude
that more radical innovations in combustor design are
required to meet EPA's emissions standards.
IV. A D V A N C E D COMBUSTOR TECHNOLOGY A N D
C O N T R O L TECHNIQUES

To emphasize that major design modifications will


be necessary for significant decreases in turbine engine
pollution, Verkamp et al. 98 show how variable air
addition (variable combustor geometry) and staged fuel
injection can keep a eombustor within the CO and NO
standards simultaneously. In Fig. 6 is reproduced a
schematic representation of how such a combustor
operates, in terms of the design values of mean primary

Gas turbine engine pollution


zone equivalence ratio and overall combustor
equivalence ratio (or engine power setting); the conventional combustors discussed in the previous section
operate along the line labeled single zone. With variable
airflow and fuel staging it should be possible to use only
the pilot flame at low powers and then shift to the main
flame at engine design conditions, as shown in Fig. 6.
1.0

O_
I,,~
r,w
o
z
w
"J 0.5

NOX LIMll
LOW
EMISSION
BAND

119

possible in principle to tailor z~l for minimum


emissions of HC, CO and N O over the LTO cycle.
Sliding bands which cover and uncover holes in the
liner provide the capability to accomplish geometry
variation continuously and remotely (see for example
refs. 6 and 98) and can be expected to introduce new
design problems involving thermal expansion, durability, and automatic control. However, the limited
data now available suggest that variable geometry
cannot meet aircraft standards unless used in conjunction with the other major modifications listed in
Table 9, 9 2 ' 9 8 i.e., unless the mixing controlled liquid
spray flame is eliminated.

CO LIMIT

2. Combustors with Staged Fuel Addition

limB" CONVENTIONAL
. . . . . . STAGED COMBUSTION ( 2 ZONES)
------ VARIABLE AIRFLOW
IDLE
MAX
OVERALL EQUIVALENCE RATIO

FIG. 6. Mean primary zone equivalence


ratio control through variable airflow and
fuel staging (from Verkamp et al.98)
These are two of EPA's control techniques requiring
major modifications (t7 and tS). All are listed in Table 9
and are expected to be developed in time for inclusion
in "next-generation" aircraft engines. Prevaporizing/
premixing combustors (control method t2) have now
enjoyed considerable testing for both aircraft and
automotive propulsion, and are discussed in section
IV.3 below.

The concept here is to replace each fuel injector in an


engine with two injectors in series (and thus with two
flame patterns). One nozzle is designed expressly to
handle the lower fuel flows near idle, and the highpower flow is split between the two injectors. In a
sense this modification is reminiscent of the design
logic for duplex pressure-atomizing nozzles, but staged
fuel addition also separates the combustion regions
and thus yields more flexibility for simultaneously
achieving good combution efficiency, stability and low
emissions. One scheme which has been used to test the
attractiveness of the control technique is shown in
Fig. 7A s6 (see below); a conceptual design also
employing premixing and prevaporization (section
IV.3) is suggested in Fig. 7B. 98
SECONDARY FUEL

TABLE9. Emissions Control Techniques Requiring Major


Combustor/'Injector Modifications2~'
EPA designation
t2-- major combustion
chamber redesign

t7--variable geometry
combustion chamber

t 8 - staged injection
combustor

Modification
Major modification of
combustion chamber and fuel
nozzle incorporating
advanced fuel injection
concepts (carburetion or
prevaporization)
Use of variable airflow
distribution to provide
independent control of
combustion zone fuel/air
ratio
Use of advanced combustor
design concept involving a
series of combustion zones
with independently
controlled fuel injection in
each

PRIMARY FUEL

(B)

1. Variable Geometry Combustors

/
PILOT COMBUSTION ZONE" ~
MAIN CMOBUSTION ZONE"-~
\
VARIABLE GEOMETRY

By redirecting the airflow split between primary and


dilution zones as the engine load is varied (via the
opening and closing of air penetration holes), it is

FIG. 7. Examples of combustors with staged fuel


injection: (A) experimental combustor of Lefebvre
and Fletcher; 56 (B) proposed configuration of
Verkamp et al. 98

JPECS Vol. 1, Nos. 2/3

120

A.M. MELLOR

Using a piloted airblast atomizer and a simplex


pressure nozzle for the primary and secondary zone
injectors, respectively (Fig. 7A), Lefebvre and
Fletcher 56 demonstrate that staged fuel addition can
reduce exhaust emissions of NO by a factor of 2 below
those produced by the liner operating in the conventional mode at the same total fuel flow. No HC
and CO results were reported, but their emissions
may increase due to incomplete combustion of the
fuel injected into the secondary zone. Mosier et al. 68
injected liquid fuel along the wall in this zone somewhat
like the scheme shown in Fig. 7B : HC and CO emissions
increased with increasing secondary fuel flow, due
particularly to quenching by film-cooling air. More

(A)

PREMIX / VAPORIZATION

3. Prevaporizing, P r e m i x i n 9 Combustors

Use of carburetion followed by either a conventional


or a catalytic combustor has also been examined. We
discuss the former first to establish the advantages of
prevaporizing and premixing the liquid fuel before
fillowing its combustion.
(a) H o m o g e n e o u s combustors
As noted several times, mean primary zone
equivalence ratio is a less significant design parameter
when dealing with diffusion flames; however, if the
liquid fuel can be vaporized and mixed with the
combustion air before it is admitted to the burner, then

AIRFLOW
CONTROL BANDS --7~.~LUTION

AIR

C.AMBER

PRIMARY
AIR

,U.L

.~

r -'1
Cz:.J~;

PRIMARY

~ I : I X I Z ~DILUTION
* |
I

/~11~/~L~ PRIMARY
/ x f l l l II/11!COMBUSTION

ZONE /
.

" '--I

DILUTION AIR
(B)

PRIMARY
AIR

DILUTION AIR

PREMIX / PREVAPORIZATION
TUBE

~1

~I

~'

FUEL
PRIMARY
AIR

UTi
FIG. 8. Examples of experimental prevaporizing/premixing combustors: (A) for small engines (from
Azelborn et al. 6) and (B) for large annular engines (from Mosier and Roberts69).

optimum injection may circumvent this unwanted


combustion inefficiency; Lefebvre and Fletcher s6 and
Mosier et al. 68 recognize that more practical designs
will be required for staged fuel injection.
N O emissions are reduced in the Lefebvr~Fletcher
burner.via staged combustion, which can be interpreted
in terms of the mean equivalence ratios for each flame.
Since more combustion air per unit mass fuel is
provided with staged fuel injection, and since less fuel
flow to the traditional primary zone is required, the
overall equivalence ratio at which combustion occurs is
more lean. Thus T,o is increased. Alternatively, the total
~sl must be less for staged fuel injection to lead to
decreased N O emissions; this occurs because there is
less fuel to mix with air in the primary zone and
because the additional airflow available in the
secondary zone flame provides increased local mixing
rates (but can also quench HC and CO reactions).
Proper design should use this decreased Ts~ to increase
combustion efficiency for this double-diffusion-flame
burner.

qSpwould become a true variable. Prevaporization and


premixing also eliminate Teb and TIi from the combustion process.
Two combustors of this type are shown in Fig. 8:
part A depicts a prevaporizing/premixing combustor
suggested for use in small gas turbine engines suitable
for automobiles and light duty vehicles 6 and part B is a
cross-section through one fuel injector in an annular
design for aircraft. 69 Note that in both combustors the
prevaporization and premixing zone is physically
separated from the primary zone (where the flame is
to be stabilized), but that both use the conventional
recirculation zone method for ignition of the premixed
fuel and air, and for flame stabilization. Thus all that
has been changed in these combustors (from the
conventional design) is that the fuel has been premixed
with the fresh air instead of with the recirculating
burned gases (compare Figs. 3C and 9). However, this
change is responsible for the considerably lower NO
emissions obtainable from such combustors: since the
inlet fuel/air mixture can only be diluted by secondary

Gas turbine engine pollution

../

///////////,/,/,/'////

FUEL + AIR

FIG. 9. Simplified schematic of the primary zone in a


homogeneous prevaporizing, premixing combustor.
air and recirculating burned gases, the feed equivalence
ratio is the m a x i m u m anywhere in the combustor.
Although turbulent heat transfer from the recirculation
zone is primarily responsible for ignition and stability,
the mean primary zone equivalence ratio at which
combustion occurs is equal to or less than that of the
fresh fuel/air mixture; thus this equivalence ratio for
combustion is unity only if the designer so chooses,
unlike the situation in a turbulent mixing (of fuel with
air) controlled, diffusion flame burner. By decreasing
this inlet equivalence ratio below stoichiometric, 3,o
can be significantly increased without too large an
increase in 3he or a degradation of combustor stability.
The result is a substantial decrease in NO with a small
but acceptable increase in HC and CO emissions.
For example, as was shown in section II, stability
requires that ~st > rh~.~, and Vranos 99 computes that
for the typical primary zone (in a prevaporizing/premixing combustor so that the is more easily evaluated,
from the feed equivalence ratio) this requirement is met
by as much as two orders of magnitude (compare
z~t* and zh~ in Table 6). Even if his numerical value of
z~z is too high, Vranos's results show that stable
operation is expected for a feed fuel/air equivalence
ratio of 0.74).8 (and undoubtedly considerably less).
For such equivalence ratios r,o >> ~ > zh~ (Table 6),
so reduction in NO without affecting stability is
possible. The key is then designing the prevaporizing/
premixer to obtain an equivalence ratio below 0.8 but
sufficiently greater than the lean flammability limitt
to preclude greatly increased HC and CO emissions
(purely from kinetic considerations; as noted previously
quenching also determines their exhaust plane concentrations). This optimum is responsible for some of the
homogeneous prevaporizing/premixing combustors
demonstrating substantially decreased emissions of
NO6,

9z

Since these combustors were also designed to reduce


idle emissions of HC and CO and inasmuch as z~b and
zi~ can determine these same emissions, improved idle
combustion efficiency should be expected and is
obtained. 6'92'69 Problems can, however, be encountered with N O emissions: Fig. 10, a comparison
of the exhaust emissions index of N O versus overall
* Unlike the situation in a conventional combustor where
Tstdetermines both turbulent heat and mass transfer, here it
refers primarily to turbulent heat transfer.
t The lean limit is further discussed in section IV.3.c.

121

equivalence ratio for the burner shown in Fig. 8B and


for the Pratt & Whitney JT9D jumbojet combustor,
shows slightly lower N O is emitted by the latter,
conventional combustor. Troth et al. 92 also found
similar NOxEI and in one prevaporizing/premixing
configuration increased NOxEI, in their experiments
with this type of retrofit combustor for the Detroit
Diesel Allison T63 helicopter engine. Such results most
likely indicate inefficient operation of the prevaporizing/premixing feed tubes. 97
30
28
26
24
22
20
=

},
ul
z
I0
8
6

o~

I
0.004

~
I
I
0.008
0.012
0.016
O V E R A L L FUEL*AIR RATIO

[
0.020

I
O.t~4

FIG. 10. Comparison of variation in measured


NOxEI with overall equivalence ratio for a prevaporizing/premixing combustor (shown in Fig. 8B)
and the conventional JT9D combustor (from Mosier
and Roberts69).

(b) H e t e r o g e n e o u s combustors
Catalytic reaction surfaces offer another method of
realizing the gains associated with premixing and
prevaporization: both Pt-coated ceramic monoliths 1o2
and catalysts of proprietary composition ~1 have
exhibited extremely low emissions u n d e r favorable inlet
conditions (NOxEI < 0.1, t/b > 99.999~; compare with
Table 3). Thus elimination of zeb, zi~ and 3s~ (and the
resulting control by zh~ and 3,o) in favor of low overall
equivalence ratio combustion, as well as the heat sink
provided by the catalyst matrix, yields the substantial
reductions in N O expected for lean, truly premixed
combustion without combustion efficiency penalties,
The flame is stabilized by the hot catalyst matrix acting
as a heat source for the incoming reactants.
Wampler et al. a2 note that their measured combustion efficiencies (at 1 atm) are in excess of those
predicted by a General Motors Research Laboratories
model which assumes mass transfer to the catalytic
surface is rate-determining, and Blazowski and
Bresowar I ~ found no effegt of pressure on qb between
4 and 11 atm. Both of these results suggest a

122

A.M. MELLOR

considerably more complicated burning mechanism


than transport-limited surface reaction.
Practical development problems obscure future
application of catalytic combustors to transportation
engines? / These include the requirement of a
minimum equivalence ratio (and thus probably variable
airflow geometry) and a minimum inlet air temperature
for good efficiency over the entire load range. Cold
starts also present a special difficulty. To attain
combustion intensity(heat release rate per unit volume)
commensurate with conventional combustors will
involve large and heavy combustors, and the durability
and lifetime of the catalysts have not been examined
thoroughly. Also, the presence of large surface areas
in the hot reaction zone increases radiant heat transfer,
which can lead to preignition of the fresh fuel/air
charge by upstream surfaces thus heated. 1~
(c) Lean limits and flashback

practical problems encountered to date with combustors of advanced design must be resolved before
such burners will find their way into commercial
engines. Examples are automatic control and durability
for variable geometry; length versus combustion
efficiency for staged fuel injection; cold starts, size,
weight, durability, and preignition for catalysts; and
lean limits and flashback for prevaporizing and premixing combustors.
We do not mean to imply that the advanced design
techniques have demonstrated across-the-board capability of meeting the emissions standards: poor combustion efficiency (high HC and CO emissions) at high
combustor loading encountered with staged fuel
addition 68 and the occasional inability of prevaporizing/premixing burners to reduce NO emissions under
those of conventional burners 69,9z demonstrate that
application of major modifications is not straightforward, primarily because understanding of the gas
turbine combustion process is not yet complete.* One
highly contributive factor here is the substantial
reliance upon measurements at the exhaust plane of
the combustor or the engine to provide information
on events within the combustor.

Certain other difficulties, in addition to proper design


of the prevaporizing/premixing section, are encountered with either type of combustor. A truly homogeneous inlet fuel/air mixture will not burn outside its
premixed flammability limit (see section II.l). Thus
Wampler et al. ~2 and Blazowski and Bresowar 1~
V. EXHAUST PLANE MEASUREMENTS
report poor combustion efficiency for overall equivalence ratios less than some minimum, and ignition
Many combustor and/or engine exhaust plane
would be impossible for still leaner mixtures. Condeterminations of HC, CO, and NO concentrations
ventional diffusion flame burners exhibit wider limits
have been cited in previous sections to help assess the
because the unmixedness (or presence of droplets)
effects of various control techniques. During 1971 EPA
guarantees that at least part of the primary zone will
commissioned several s t u d i e s 2'39'96,71 to ascertain
have a composition adequate for combustion.53
baseline emissions of engines currently in use. Most of
However, in aerodynamically stabilized premixed
these data were subsequently analyzed 14.5s,s 3 and used
combustors the premixed limits are anticipated to
to determine both EPA standards z6,27,47 and USAF
apply, because local equivalence ratios are at most
goals. 12'13'4I Our intent here is simply to determine
equal to the inlet value, a result of mixing with the
to what extent such data yield information on the
(fuel lean) recirculating gases.
combustion process.
The propagation of flame from the primary zone
upstream through the premixer is referred to as flash1. Correlations of Exhaust Emissions
back and generally will result in prevaporizer
burnout. 11'1z Flashback occurs (due to the presence
Lipfert 58 was the first to correlate engine exhaust
of fuel in the feed stream) if the approach velocity
concentrations, primarily in terms of engine cycle
everywhere decreases below the premixed flame speed
design parameters, in order to determine acceptable
(compare Figs. 3C and 9). It can be avoided if the flow
levels of data scatter and the influence of second order
time between fuel injector and flame stabilizer is
effects such as relative humidity of the engine inlet air.
maintained smaller than rhc evaluated under these
Two examples are shown in Figs. 11 and 12: Fig. 11
flow conditions (see also the discussion of flame stability
gives EI HC and E1 CO versus combustion inefficiency
in section |I.2).* Unfortunately, as is also true for
(1 --t/b) for the Pratt & Whitney JT8D and JT9D and
flammability limits, flashback or flame speed data are
Rolls Royce Spey engines, based on data reported by
not available for fuels of interest at the high turbulence
Bogdan and McAdams; 14 NOxEI for several engines
levels and elevated temperatures and pressures of the
is correlated with combustor inlet air temperature in
compressor discharge, and design must proceed on a
Fig. 12. From perfectly stirred reactor calculations
trial and error basis, x~
compared with Fig. 12, Lipfert 5s concludes that a
mean primary zone equivalence ratio of unity and a
4. Summary
mean primary zone "residence time", constant for all of
Combinations of EPA's major modification control
the engines included in the correlation, could explain
techniques can meet aircraft standards, but the many
the universality of the curve.
* Due to lower combustor loadings, catalytic combustors
are more susceptible to flashback,

* Although empiricism has eliminated smoke, NO is much


more difficult to affect..

Gas turbine engine pollution


I00

123

Zeldovich mechanism and local equilibrium of O and


02 (see ref. 15). O n this basis these investigators
conclude that the "characteristic time for N O
formation" is the same for all of the engines listed in
Fig. 13.
Note that the assumption of stoichiometric combustion by both Lipfert s8 and Sawyer et al. 83 implies
mixing-controlled burning in the primary zone; why
the time factor should be of the same order from
engine to engine is suggested by a physical model 64
of the turbine combustion process, also developed from
exhaust plane data.

CO

IO

40
30

O
t3

20

I0

o1,
.I

.001

.01
BURNER INEFFICIENCY ( I"11B)

"'

LIEGE
JT3D
~
JT 8D

J T 9D

R.R. SPEY
o
J52
I~
J 57
,~
T56
o
TF 30
-eOLYMPUS 5 9 3 1

0.I

FIG. l l. Correlation of HCEI and CO EI with


combustion inefficiency (from Lipfert58).
Sawyer et al. 83 extend the NOxEI correlation to
several additional engines, but prefer to use the
stoichiometric adiabatic flame temperature (based on
inlet air temperature) in a more traditional Arrheniustype graph shown in Fig. 13. The straight line in this
figure corresponds to an activation energy for NO
formation of 136 kcal/mole, which is consistent with the

40

3.5

,o~
~.\
\
\
\
\
[

4.0
I
X I0 4 ( K -~)
Tmox.

4.5

FIG. 13. Correlation of NOxEI with


stoichiometric adiabatic flame temperature, based on combustor inlet air temperature (from Sawyer et a/.83).

ENGINE MODEL

I
/

:50

2. A Physical Model Jor Conventional Combustors


Suggested by Exhaust Emissions

GTC85-90-2

~o-l 2 0

o~
Z

200

400
COMBUSTOR

&,

600
8
iNLET TEMPERATURE,*F

&

FIG. 12. Correlation of NO~EI with combustor inlet air temperature (from Lipfert58).

O n the basis of variations of NOxEI and CO El,


both with combustor reference velocity (V~ef)in a liner
of the penetration jet type, and with differential fuel
injection pressure (Ap) across its simplex pressure
atomizing nozzle, Mellor 64 postulated the two-part
physical model shown in Fig. 14 for spray combustion
in a conventional burner. High values of NOxEI and
CO EI at engine idle conditions (Fig. 14A) are
attributed to long ~eb (either slow droplet combustion
or evaporation), whereas as engine power is increased
(Fig. 14B and 14C) "~eb becomes small compared to
zs~, so that the combustion process resembles a
turbulent diffusion flame (with an imbedded recirculation zone). Here NOxEI is observed to be independent
of Ap and CO E1 increases with Ap. Since the air
flowrate (or reference velocity) determines Ts~,constant
NOxEI results from constant zst: the conclusions of

124

A.M. MELLOR
Drop Diffusion Flames
yellow Flame
Long Flame Length
NO High Since Drop
Flames StoichiornetrK:
CO High Since rob Long,
Incomplete Combustion

Ca)
o * Large ~eb

(b)

t
~=p~15Otto

Transition from Drop


Diffusion Flames to
Woke Flames or Pure
Evaporation

Minimum NO and CO

Drops Evaporate

Blue Flame

(c)

Turbulent Diffusion

Small feb

Flame of Constant
Length
NO~f(.~o~ Since Flame
Length Constant
CO ~(~p)Since ~o ~P
and Flame Length
Constant
-~

FIG. 14. Schematic representation of physical model


of spray combustion in a conventional primary zone
with a simplex pressure atomizing nozzle: (A) engine
idle, (B) midrange and (C) engine cruise (from
Mellor64).
Lipfert 58 and Sawyer et al. 83 that the engines listed in
Figs. 12 and 13 have very similar "residence times,"
or "characteristic times of N O formation" are thus
reinterpreted in terms of very similar mixing times in
the shear layers surrounding the flameholding recirculation zones in the various (conventional) engines.
C O EI is high for this particular liner at full power
due to quenching by the jet cooling. 64
The physical model of the combustion process as a
function of engine loading (Fig. 14) should remain at
least qualitatively correct for conventional combustors
with other types of fuel injectors. For example, duplex
pressure atomizers exhibit a different functional
dependence of Sauter mean diameter on fuel flowrate, 73
and therefore Zeb is not simply related to Ap.
If air-assist and airblast atomizers (section III.1) do
in fact produce finer atomization than their pressureatomizing counterparts, then feb appropriate for such
nozzles will be smaller. However, since air is now
introduced with the injected fuel rsi may replace feb
in determining low power emissions: changes in air
injection pressure may have a more important effect on
local mixing and combustion than on droplet size
distribution, at least as far as idle emissions are
concerned. Mellor 65 has attempted to correlate N O
emissions from a kerosene spray flame fueled by an
air-assist nozzle 5 in terms ofzeb and the physical model
shown in Fig. 14, by arguing that the experimental
characteristic time of N O formation is proportional
to the mean fuel droplet lifetime. The analysis was
prompted by the inability of a purely mixing controlled theory to explain the dependence of N O
emissions on Apa, the air atomizing pressure drop
across the injector. 5

However, the reported Apa on which both analyses


were based were subsequently found in error, 4z and
when corrected do agree with the mixing controlled
model. ~9 rfi, evaluated via the technique of Corrsin 23
(see Vranos's 99 analysis cited in section II.2), is the
characteristic time for N O formation. This new'result
has been verified by injecting gaseous propane rather
than liquid kerosene: the same numerical value of zfi
is deduced from either set of data. When the air-assist
nozzle is replaced by a simplex pressure atomizer
injecting kerosene, however, the experimentally determined characteristic time of N O formation correlates
with zeh rather t h a n "~fi,49 consistent with the results
obtained from the Allison J-33 combustor and
injector. 64 Therefore the larger of zj-i and Zeb can
control idle emissions from conventional combustors
(section III.1), and the physical model of spray combustion at low engine power settings (Fig. 14A) should
be generalized accordingly.* r~t is determining at higher
combustor loadings, as indicated by the model.
Baseline emissions and qualitative physical models
for gas turbine combustion can thus be obtained from
burner or engine exhaust measurements. However,
the details of this complex combustion, especially useful
for interpreting local quantitative effects of combustor
design changes and for comparison with analytical/
numerical models for calculating emissions (section
VII), are only available if gas sampling and other fluid
properties (e.g., temperature, velocity magnitude and
direction) are obtained as a function of position within
the burner. 63

VI. INTERNAL MEASUREMENTS

The inadequacy of exhaust measurements was noted


as early as 1956 by Cornelius and co-workers. 22 In a
study to find burner designs which would reduce smoke
and odor emissions from the Allison T56 engine, they
obtained from gas sampling and temperature measurements within the combustor contour maps of CO2, CO,
02 concentrations, local equivalence ratios, smoke
numbers and gas temperature estimates, in a test facility
capable of simulating compressor discharge conditions
for both ground idle and military power settings.
Typical temperature contours for three liner designs
are shown in Fig. 15. F r o m such data combined with
combustor exhaust plane concentrations (which
include formaldehyde), very specific liner design
modifications were suggested. 22 Toone 91 reported
results of a similar study on smoke.
Interest in internal measurements later revived due
* Liquid propane was the fuel used in the study which
led to the development of Fig. 14, and it now appears that at
some of the combustor loadings investigated a portion of
the fuel vaporized within the simplex injector. 95 Whether
the presence of vapor influences zebor r/iis presently unclear;
poorer atomization (larger Zeb)due to collapse of the spray
cone at low differential injection pressures seems more
likely, since in other experiments replacing liquid by gaseous
fuel usually results in lower, not increased, NO
emissions.V9,72.85

Gas turbine engine pollution


OPERATING CONDITION - Militory Power(~ Seo Level
FUEL NOZZLE - DPD X-2t155
FUEL - J P ' 5
Stgtion S ~
St(Itl~n S1ot~ll

~'~
2500

~-

SlotJon

i?.000
I000
L w v 2500

IPNI~ED~RCTION
IOOO

'

2 O0

~2"2800

2000

TOROIOAL
LINER

FIG. 15. Temperature (F) contours in three Allison


T56 combustors at military power (JP5 fuel) (from
Cornelius et al.22).
to lack of information on the NO formation process
within liners. Workers at Berkeley 84'89,v9 gas sampled
in a laboratory model combustor at ambient inlet
temperature and slightly greater than atmospheric
pressure. NO contours for three different overall

125

equivalence ratios (varied by changing air flowrate at


constant fuel ftowrate) are presented in Fig. 16; CO2,
CO and HC, and temperature and local overall
equivalence ratio inferred therefrom are also
reported. 89 Highest local NO concentrations correspond well to peak temperature, slightly lean zones.
To explain lower NO emissions with methanol
rather than # 2 diesel oil from an automotive combustor, LaPointe and Schultz 5'5~ report analogous
results obtained in a high-pressure, heated air facility.
Data for other types of combustors are now becoming
available a s w e l l . 9 0 ' 1 7 ' 3 4 ' 6 1 ' 9 9
Local concentrations and gas temperature estimates
have also been employed to develop flow models for
both penetration-jet66,93 and film-cooled combustors, s5 Figure 17 shows results for the former, and
in Fig. 18 various regions are identified for the Allison
T56 liner (compare Fig. 15): region 1 is the hollow
cone fuel spray which feeds the turbulent flame
extending through regions 2 and 3 and closing to the
centerline in region 5. NO, CO and HC are all high
in region 2, with substantial further reactions beginning
in region 5 (corresponding well with the maximum
temperature zone reported by Cornelius e t al. z z for the
modified production liner in Fig. 15) and continuing
into region 6. Region 4 is the centerline recirculation
zone characterized by relatively constant concentrations of NO, CO, and HC. 85
Internal measurements for conventional combustors
and the derived flow models, particularly those
obtained at high engine power settings, are thus

03

~l.O
-0.ll9

O0

I0

12

II

13

14

1.5
~'l.O
z

I J I I.-oioo
~ ~"~

~2

I0

1.5

.5
J

12

~r~

l.O

~o

II

13

~'~

Iw,,

14
WALL

-o..
~-0.281
8

I0

II

12

13

AX,S

14

AXIAL DISTANCE, INCHES

FIG. 16. NO (ppm) contours at three different overall equivalence ratios for a laboratory
model combustor (1.5atm, 300K inlet air temperature, n-heptane fuel) (from Starkman
et

a/.89).

126

A.M. MELLOR
I. Very fuel rich; much CO
I formation; fuel drops
/
continue downstream
/
~

2. CO quenched by cold
3. Measured temperatures low
recirculoting air NO
/
due to drop impingement;
entrained from wall flow . / J
some CO oxidation and NO
~
formation
- - ~

ii _d75

/. . . . . . . .

7.,-

- > < .

li_gw_4

I m

I~',\'~\

4. Penetration air

_ .

~.'x"~ ........

zo.

5NO formation

~ i~,~/

~ i ~ l /

.........

high temperature

region 3-5

i I

5\ ,'_if Z

17.2
Distance from injector,cm

Fro. 17. Postulated mean flow model for the primary zone of the Allison J33 penetration-jet
combustor (5atm, 274 or 500K inlet air temperature, 2.04 or 2.72kg/sec airflow, overall
equivalence ratio about 0.2, liquid propane fuel) (from Mellor et al. 6~ and Tuttle et al.93).

generally quite consistent with the physical model


shown in Fig. 14C. For conditions where r,,~, < %t,
mixing between the hot burned gas/fuel vapor
aggregate in or near the recirculation zone and the fresh
air admitted through the dome and liner (regions 2, 3
and 5 in Fig. 18) determines the rate of fuel combustion

70

RADIAL
POINT
4
3,5
2
6

exhaust plane measurements, and supported by the


available internal measurements, can be used to probe
weaknesses of the available calculational models of the
gas turbine combustion process. Gas sampling within
combustors also provides data for comparison with
their output.

-,
v

-i '''~--k - ~ - - J ~
"---I \
I ~, !
I
'C
i

_J-~"

i
,

i1-~

'.dT/z/~k"

I ~
I
I~

',--'i

I
!

"
-!

'

3.28 5.82 7.65

13.36
17.98 21.23 24.492733
Axial Planes (cm. from injector)

x;

32.74 36.42

FIG. 18. Postulated mean flow model for the Allison T56 film-cooled combustor (9.6atm,
584K inlet air temperature. 2.4kg/sec airflow, overall equivalence ratio 0.299, liquid
propane fuel). Numbered regions are discussed in text (from Shisler et al.SS).

and N O formation; the situation rather resembles a


turbulent diffusion flame surrounding a reversed flow
zone (region 4). Laboratory studies of spray combustion
have also generated similar findings. 19'2'2x'62-7(' The
implication is that the spray is fed by rapidly vaporizing
liquid fuel and tends to burn as a whole. However,
according to the physical model, for cases where
atomization is poor or the fuel is of low volatility
(i.e., where r~b > zsl), slow droplet evaporation controls
the combustion: the possibility of burning also occurring via droplet diffusion flames or wake flames cannot
be ignored. 64'86 Mixing in the immediate vicinity of
the injector introduces a third characteristic time zyi,
which can become important if r/i > z~b. As suggested
in Fig. 14, in conventional gas turbine engines the rate
of heat release is expected to be controlled by rf~ or
feb only at low engine powers.
This physical model, suggested by combustor

Vll. A N A L Y T I C A L M O D E L S

Combustor modeling has been reviewed recently by a


number of a u t h o r s . 36"63"8A6,78,75 We limit our
attention here to models which are capable of
predicting (or correlating) pollutant exhaust emissions
or species concentrations within liners and which have
been compared with measured data; see Odgers 75 and
H a m m o n d and Mellor 36 for historical reviews,
particularly of global, performance-oriented models.

1. M o d u l a r

Models

Those models which consist of chemical reactors of


various types (e.g., partially or well-stirred reactors)
in series and/or parallel, designed to simulate largescale regions of the conventional combustor, are

Gas turbine engine pollution

AIR

SECONDARY
ZON_E_

PRIMARY ZONE

R fu

Rsi

- -

127

SECONDARY
~ONE

DILUTION
ZONE

1 ii1

R2
FUEL

"1

FIG. 19. Overall flow model schematic of Hammond and Mellor. 37


referred to as modular. Calculations from three of these
have now been compared with experimental data.
One 37 is based on the combination of perfectly
stirred reactors (R1, R2, and V~=) shown in Fig. 19.

A second model, developed by Fletcher and


Heywood, 29 has been extensively tested against N O
exhaust emissions from both conventional, 3 advanced
aircraft, 44'56 and automotive 44 combustors. Depicted

102:
A EXPERIMENTAL DATA OF REE(66)

n
,j
z

R2

I0
RI

o
z

c.
10
0.CO

,dco 2010o 30'.0o ~bo

5o'.co 6o.'oo ro'.co 8o.'oo 9o'.0o

AXIAL DISTANCE FROM START OF FLAME TUBE (CM)

FIG. 20. Computed and experimental mass-averaged axial NO profiles in the


Allison J33 combustor; the two computed curves are for differing perfectly
stirred reactor volumes downstream of the primary zone (see Fig. 19) (from
Hammond and Mellor38).

Droplets are ignored entirely, and a recommended 63


homogeneous chemical kinetic mechanism based on a
global, finite rate, hydrocarbon partial oxidation
obtained empirically by Edleman and Fortune 25 is
used. Stirred reactor volumes, flow splits, etc., are
selected to simultaneously yield the best fits with
internal NO, CO, and temperature measurements as
reported 6 for the Allison J33 combustor; this fit for
the mass-averaged axial N O profile is shown in Fig. 20.
Other results indicate that good agreement cannot be
found for one set of parameters if both N O and CO
profiles are to be fitted and that the model predicts
very poorly the quantitative variations in N O and CO
exhaust emissions with changes in combustor loading. 38 These difficulties result primarily from both
neglect of heterogeneous effects (see section V.2) and
modeling a turbulent diffusion flame with an imbedded
recirculation zone as two large interconnected perfectly
stirred reactors in parallel (compare Figs. 14C and 19).

schematically in Fig. 21, the primary zone is modeled


in part as a partially stirred reactor (PSR) burning
gaseous fuel over a Gaussian distribution f ( ~ ) about
the mean primary zone equivalence ratio; a downstream lateral mixing reactor narrows this distribution
to the mean and is followed by a plug flow reactor.
AIR

FUEL

PSR

ID

VAPOR
LATERAL
MIXING

AIR

T, f (,)

FIG. 21. Primary zone flow model schematic of Fletcher and Heywood z9 (from
Mellor63).

128

A.M. MELLOR

Reactor volumes, airflow distribution, and the standard


deviation of the equivalence ratio distribution are
obtained from the best fit of model output with
measured exhaust plane N O emissions as a function of
combustor loading, where N O is calculated from the
modified Zeldovich mechanism assuming local C - H - O
equilibrium. 29'63. Good correlation is usually obtained
except at low fuel/air ratios (engine idle), as shown in
Figs. 22 and 23 (s is a mixing parameter related to the
width of the 4) distribution; for s = 0 the reactor

~//~

s - O3 3
O. 42

0
z

o
o
u~ I0 -4

600
COMBUSTOR

400

- E

i0 -3

A
B

w"
C3

CORRELATION FOR

~- =

0
o_
n,I.-

1.15 x CORR.

z-2o~
o

[.

f-

0.42110.33

t100
z

80

COMBUSTOR
INLET TEMPERATURE
=900F(NOMINAL)

6O

10-5

g 4o
u

0.01

0.02

0.03

0.04

0.05

0.06

0.07

FUEL- AIR RATIO


IC
40

60

80

I00

120

140

FIG. 23. Computed and experimental massaveraged exhaust emissions of NO versus fuel/air
ratio for the NASA swirl can combustor; s is a
model fitting parameter (from Heywood and
Mikus44).

I
160

AIR FUEL RATIO

F[o. 22. Computed and experimental mass-averaged


exhaust emissions of NO versus air/fuel ratio for two
aircraft combustors; s and ~ are model-fitting parameters (from Fletcher et al.3).
becomes perfectly stirred). In the first figure the model
extrapolated to idle works poorly for combustor A, and
in Fig. 23 two mixing scales are required for matching
the data at low and high engine power for this
carbureted combustor.
If the Gaussian distribution function is in reality
curvefitting a concentration fluctuation frequency
distribution in a turbulent diffusion flame ~ rather
than an equivalence ratio distribution in the primary
zone, then the success of the Fletcher-Heywood model
at high engine loadings is not unexpected: 64 the
partially stirred reactor is a simplified calculational tool
for N O formation in the mixing-controlled flame of
Fig. 14C (see also ref. 52). At low fuel/air ratios the
discrepancy for combustor A could be a result of
ignoring feb, and the change in scale required for the
swirl can combustor (Fig. 23) is entirely consistent with
TI~ determining N O emissions at idle.
The third modular model (Fig. 24) includes both T~l
and T~b, at least conceptually. Here streamtubes are
allowed to exchange mass and heat by turbulent mixing,
and penetration and film-cooling air injection is
modeled semi-empirically. 68'6~ Droplet burning is

allowed in streamtubes 2 and 3 by assuming stoichiometric combustion of fuel vapor added to the flow at a
rate corresponding to the convection-enhanced "d 2
law". An extended N O / N 2 0 mechanism is used for N O
formation 69 and a fuel partial oxidation mechanism
based on the global finite-rate expression of Edelman
and Fortune 2~ is followed by a rate-constrained partial
equilibrium C O mechanism. 61'7 Considerable experience with model predictions has allowed best
selection of its arbitrary parameters for conventional
combustors; for example, it is found convenient to use
in the streamtube mixing analysis a local eddy viscosity
empirically adjusted to be a function of burner inlet
air temperature (thus implicitly a function of combustor
loading). 7
COMBUSTION

* This assumption precludes meaningful CO and HC


emissions predictions.

DILUTION AIR- ~ . /

FUEL ~OZ--'~LE

.[

r COOLING AIR

~-"

Flo. 24. Overall flow model schematic of Mosier


et al. e8

Gas turbine engine pollution


I0 2

129

2. C o n t i n u u m M o d e l s

.J

~lo
CO

-\ \ \\

NO x

CLOSED SYMBOLS- PREDICTED


OPEN SYMBOLS - MEASURED
tO (

i0"1

i(# 2

400
BURNER

600
INLET

BOO

I
- - I000

[
1200

I
1400

TEMPERATURE ~ ' F

FIG. 25. Computed and experimental mass-averaged


exhaust emissionsversus burner inlet temperature for
the Pratt & Whitney JT9D combustor (from Mosier
and Roberts7).
Comparisons of predicted and measured exhaust
emissions of HC, CO and NO are quite satisfactory
for most of the conventional combustors examined to
date: the Pratt & Whitney JT9D and Allison T56 results
are shown in Figs. 25 and 26, respectively. Since the
model is by far the most complete currently available
in terms of including all of the relevant processes
indicated in Fig. 14, the agreement is most encouraging.
Examination of its solutions inside the combustor in
light of presently available internal measurements (Fig.
18) will allow its further refinement (see also ref. 61).
I02

.j I0 I
u.

If the number of streamtubes used in the MosierRoberts analysis were greatly increased, then the model
could be considered of the continuum type as discussed
by Gosman et al., 32 Spalding,s8 Caretto, 16 and others;
here the reactor or computational grid dimensions are
small compared with those of the combustor. Such
models may be expected to have fewer arbitrary parameters and thus be more predictive than their modular
counterparts, but the requisite inclusion of two-phase
and kinetic effects, coupled with the already complex
and time-consuming computer codes, has frustrated
many workers (see, for example, Hunter et al.46).
Anasoulis et al. 4 have succeeded in obtaining axisymmetric solutions of the time-averaged NavierStokes equations, including finite rate NO formation
chemistry in the manner of Fletcher and Heywood 29
and heterogeneous effects following Mosier et al. 68 To
speed the calculations a field relaxation technique and
a simple turbulence model are employed. Evaluation
of the results against measurements has been limited to
date to combustors where three-dimensional effects
may be important and thus are far from encouraging.
Temperature contours and centerline CO2 and NO
profiles within the Berkeley combustor (cf. Fig. 16) are
presented in Figs. 27 and 28, and comparison with
exit plane measurements for the annular combustor
tested by Greene and Vranos 34 also reveal substantial
discrepancies# Although the authors identify threedimensional fluid flow as the culprit, contributing
errors undoubtedly exist in their submodels for
turbulence, chemistry and droplet evaporation. However, in finding solutions and subjecting same to
examination with measurements, these investigators
demonstrate commendable fortitude. Too often only
numerical results of these calculations have been
reported, with no quantitative comparison with
measured data; such emasculated exercises are of
questionable utility.
Note that both the modular and continuum models
are generally limited to conventional burners and even
for these are far from proving their merit as design tools,
with the possible exception of that of Mosier and
Roberts.

JmlO0
x

VIII. S U M M A R Y A N D C O N C L U S I O N S

O
[]
A

CO
UHC
NO X

CLOSED SYMBOLS- PREDICTED


OPEN SYMBOLS- MEASURED

10-20

I
200

I
I
[
BOO
400
600
BURNER INLET TEMPERATURE

I
I000
~ "F

I
1200

I
1400

FIG. 26. Computed and experimental mass-averaged


exhaust emissionsversus burner inlet temperature for
the Allison T56 combustor (from Mosier and
RobertsT).

Within the context of a two-part (engine idle and


cruise) physical model of spray combustion in a conventional gas turbine engine (Fig. 14), five characteristic
times (Table 4) are identified, following many other
investigators, to determine the controlling processes for
combustion efficiency, stability, and emissions. For
most conventional combustors designed to date, the
droplet lifetime ~eb and eddy lifetimes near the fuel
injector (7ii) and in the shear layer surrounding the
centerline recirculation zone (%t)(Fig. 3) are much more
relevant than the appropriate homogeneous kinetic
times.

130

A.M. MELLOR
O

1080" R

2160" R

MEASUREMENTS

2880"R

FOR OVERALL EQUIVALENCE RATIO OF 0.205

OF STARKMAN ET AL.(REE 8 9 )

3240 =R
PREDICTION

IIIIIIIIILIIIIIIIIIILIIIIIIII[[I~[~IIIIILIII
llIIl

IIIIJJIIHIIIHIIlII~IIllIlllllllIllllIll
lllIll

1.4

I
I

O
1.2

r7
1.0

0.8

0"6 1

0.4

0.2

o
o

O 0

0 ,,
0.5

1.0

I .5

2.0

2.5

3.0

35

4.0

Flo. 27. Computed and experimental temperature contours for the Berkeley
can combustor (from Anasoulis et al.4).
In essence, the physical model suggests that under
and near conventional engine design conditions
T~ > T~b(or TI~),so that the shear layer is the dominant
source of NO:* slow mixing and thus extended gasphase combustion (long ~t) are expected to yield high
NO emissions. The effect of ~z on HC and CO is not
OVERALL EOUIVALENCE RATIO 0.205
CENTERLINE MEASUREMENTS

I000000

I00000

CO2
o

I0000

~
I000

1-

o/57-

I00

NO

o
i

'o

,'o

2'.o
3'.0
4:o
AX,AL~STA.CE.,~:.ES

b.O

FIG. 28. Computed (solid lines) and experimental


centerline axial profiles of C O 2 and NO for the
Berkeley can combustor (from Anasoulis et al.4).

* This is usually the only region within a combustor


where the local mean equivalence ratio is near unity ~md
thus flame temperatures are high enough to produce NO.

as predictable, because their emissions are also determined by quenching in, for example, the recirculation
zone or along the combustor walls by film cooling air.
Generally, however, long T~ is also anticipated to lower
the combustion efficiency.
As conventional combustor loading is decreased
toward engine idle, then the situation can reverse such
that T~b(or vSi) > Tsl.lfzeb is appropriate, then efficiency
and NO are determined by the rate of addition of fuel
to the vapor phase and under certain conditions (low
relative velocity between fuel drops and ambient gas,
fuels of low volatility, and poor fuel atomization)
droplet diffusion or wake flames are possible. ~s~ determining means rate-controlling mixing of fuel vapor
eddies with the recirculation gases. Also, when air
momentum is used to assist in fuel injection, smallscale fuel/air mixing near the nozzle (Tsi) may override
v~'s importance, since the resulting finer atomization
leads to reduced spray penetration, and since more air
is available in this region of the burner.
This physical model is supported by the available
exhaust plane emissions trends and by concentration
and temperature measurements made within conventional combustors; the analytical model of Mosier
and Roberts, which has enjoyed the most success to date
in correlating exhaust emissions of HC, CO and NO
from diffusion flame burners, includes all of these
processes via state-of-the-art calculational procedures
and assumptions.
It is extremely important to recognize that some of
the model's features and implications for emissions
reductions apply equally well to advanced combustors
involving variable geometry or staged fuel injection
techniques, which imply multiple diffusion flames in
designs proposed to date. The key to significant NO
emissions decreases thus may be accomplishing fuel/air
premixing in a zone separate from the combustion
process so that primary zone equivalence ratio is a true
design variable and does reflect the mixture ratio at

Gas turbine engine pollution


which combustion occurs, unlike the diffusion flame
configuration. Flame stabilization in such a prevaporizing/premixing burner can be obtained by turbulent
heat transfer from either a recirculation zone (homogeneous combustion) or a hot catalytic surface (heterogeneous combustion). Leaning out the equivalence
ratio then increases r,o without an overly adverse effect
on Zhc, and zeb, ryi and zsl are completely eliminated
from determining the final exhaust emissions of NO,
H C and CO.
Minor burner modifications, directed primarily at
affecting only one of these characteristic times, do not
appear capable of meeting EPA aircraft emissions
standards. Instead, advanced combustors which accomplish staged combustion responsive to engine
loading (via staged fuel injection or variable aerodynamic flow patterns) or which utilize prevaporization
and premixing will be necessary to satisfy simultaneously the traditional design criteria of highly stable
and efficient burning and that of low emissions.

Acknowledgements

The author gratefully recognizes critical


evaluations of this manuscript by Prof. Arthur Lefebvre of
Cranfield Institute of Technology, Dr. Richard Munt of the
U.S. Environmental Protection Agency and Messrs. Jack
Vaught and Frank Verkamp of Detroit Diesel Allison
Division, General Motors Corporation. Financial support
for his research on gas turbine emissions is provided in part
by grants from the Environmental Protection Agency and
from both the Research Laboratories and Allison Division of
General Motors. The contents of this paper do not
necessarily reflect the views and policies of the Environmental Protection Agency or of General Motors Corporation, nor does mention of trade names or commercial
products constitute endorsement or recommendation for
use.
REFERENCES

1. ADVISORY GROUP FOR AEROSPACE RESEARCH AND


DEVELOPMENT. Atmospheric Pollution by Aircraft
Engines, AGARD CP No. 125 (1973).
2. AIRESEARCH MANUFACTURING CO. OF ARIZ. Exhaust
emissions test: AiResearch aircraft propulsion and
auxiliary power gas turbine engines. GT-8747-R (1971).
3. AETFNKIRCH,R. A. and MELLOR, A. M. Emissions and
performance of continuous flow combustors, Fifteenth
Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, in press (1975).
4. ANASOULIS,R. F., McDONALD, n . and BUGGELIN, R.
C. Development of a combustor flow analysis. Part I:
Theoretical studies. AFAPL-TR-73-98, Part I (1974).
5. APPLETON, J. P. and HEYWOOD, J. B. The effects of
imperfect fuel-air mixing in a burner on NO formation
from nitrogen in the air and the fuel, Fourteenth
Symposium (International) on Combustion, p. 777. The
Combustion Institute, Pittsburgh (1973).
6. AZELBORN, N. A., WADE, W. R., SECORD, J. R. and
MCLEAN, A. F. Low emissions combustion for the
regenerative gas turbine. Part 2: Experimental techniques, results, and assessment. ASME Paper No.
73-GT-12 (1973).
7. BAHR, D. W. Technology for the reduction of aircraft
turbine engine exhaust emissions, Atmospheric Pollution
by Aircraft Engines, AGARD CP No. 125 (1973).
8. BARR~RE,M. Modelisation des foyers de turboreaetor
en vue de l'etude de la pollution, Atmospheric Pollution
by Aircraft Engines, AGARD CP No. 125 (1973).
9. BAY AREAAIR POLLUTION CONTROL DISTRICT. Aviation
Effect on Air Quality m the Bay Region (1971).

131

10. BECKER, K. H. and SCHURATH, U. Photo-oxidation of


aircraft engine emissions at low and high altitudes,
Atmospheric Pollution by Aircraft Engines, AGARD CP
No. 125 (1973).
11. BLAZOWSKI,W. S. and BRESOWAR, G. E. Preliminary
study of the catalytic combustor concept as applied to
aircraft gas turbine engines. AFAPL-TR-74-32 (1974).
12. BLAZOWSK1,W. S. and HENDERSON, R. E. Assessment
of pollutant measurement and control technology and
development of pollution reduction goals for military
aircraft engines. AFAPL-TR-72-102 (1972).
13. BLAZOWSKI, W. S. and HENDERSON, R. E. Aircraft
exhaust pollution and its effect on the U.S. Air Force.
AFAPL-TR-74-64 (1974).
14. BOGDAN, L. and MCADAMS,H. T. Analysis of aircraft
exhaust emission measurements. Cornell Aero. Lab.
NA-5007-K-1 (1971).
15. BOWMAN, C. T. Kinetics of pollutant formation and
destruction in combustion, Progress in Energy and
Combustion Science (Chigier, N. A., Ed.), Pergamon,
Oxford, in press (1975).
16. CARETTO, L. S. Modeling pollutant formation in combustion processes, Fourteenth Symposium (International)
on Combustion, p. 803. The Combustion Institute,
Pittsburgh (1973).
17. CASCl, C, COGHE, A., GHEZZI, U. and PASINI, S. An
experimental research on the behavior of a continuous
flow combustion chamber, Atmospheric Pollution by
Aircraft Engines, AGARD CP No. 125 (1973).
18. CERNANSKY,N. P. and SAWYER, R. F. NO and NO2
formation in a turbulent hydrocarbon/air diffusion
flame, Fifteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, in press
(1975).
19. CHIGIER, N. A., MAKEPEACE,R. W. and MCCREATH,
C. G. Aerodynamic interaction between burning sprays
and recirculation zones, Combustion Institute European
Symposium 1973, p. 577. Academic Press, New York
(1973).
20. CHIGIER, N. A. and MCCREATH, C. G. Combustion of
droplets in sprays, Acta Astronautica 1, 687 (1974).
21. CHIGIER, N. A., MCCREATH, C. G. and MAKEPEACE,R.
W. Dynamics of droplets in burning and isothermal
kerosene sprays, Combust. Flame 23, 11 (1974).
22. CORNELIUS,W., BURWELL,W. G. and TURUNEN, W. A.
Progress report on fundamental gas turbine combustion
studies. General Motors Research Memoranda (1957).
23. CORRSIN, S. Simple theory of an idealized turbulent
mixer, AIChE J. 3, 329 (1957).
24. Dtx, D. M. and BASTRESS,E. K. Approaches to the design
of low-emission gas-turbine combustion chambers. SAE
Paper No. 720728 (1972).
25. EDELMAN, R. B. and FORTUNE, O. F. A quasi-global
chemical kinetic model for the finite rate combustion of
hydrocarbon fuels with application to turbulent burning
and mixing in hypersonic engines and nozzles. AIAA
Paper No. 69-86 (1969).
26. ENVIRONMENTAL PROTECTION AGENCY. Aircraft emissions: impact on air quality and feasibility of control
(1972).
27. ENVIRONMENTALPROTECTION AGENCY. Control of air
pollution from aircraft and aircraft engines, U.S. Federal
Register 38, (136), Part II, July 17 (1973).
28. ENVIRONMENTAL PROTECTION AGENCY. Supersonic
aircraft : control of air pollution, U.S. Federal Register
39, (141), Part I, July 22 (1974).
29. FLETCHER,R. S. and HEYWOOD,J. B. A model for nitric
oxide emissions from aircraft gas turbine engines. AIAA
Paper No. 71-123 (1971).
30. FLETCHER,R. S., SIEGEL,R. D. and BASTRESS,E. K. The
control of oxides of nitrogen emissions from aircraft gas
turbine engines. I. Program description and results.
FAA-RD-71-111,1 (1971).
31. GEORGE, R. E. and BURL1N, R. M. Air Pollution from

132

32.

33.
34.
35.

36.
37.
38.
39.

40.

41.

42.
43.

44.

45.

46.

47.
48.

49.
50.

51.

A.M. MELLOR

Commercial Jet Aircraft in Los Angeles County, Air


Pollution Control District, County of Los Angeles
(1960).
GOSMAN, A. D., PUN, W. M., RUNCHAL, A. K.,
SPALDING, O. B. and WOLFSTEIN, M. Heat and Mass
Tran~/~,r in Recirculating Flows, Academic Press, New
York (19691.
GOULDIN, F. C. Controlling emissions from gas
turbines the importance of chemical kinetics and
turbulent mixing, Comb. Sci. Tech. 7, 33 (1973).
GREENE, W. and VRANOS, A. Development of a combustor flow analysis. Part II: experimental studies.
AFAPL-TR-73-98, Part II (1974).
GROBMAN,J. S. Effect of operating variables on pollutant
emissions from aircraft turbine engine combustors,
Emissions,]rom Continuous Combustion Systems, p. 279.
Plenum Press, New York (1972).
HAMMOND,D. C., Jr. and MELLOR, A. M. A preliminary
investigation of gas turbine combustor modelling,
Comb. Sci. Tech. 2, 67 (1970).
HAMMOND, D. C., Jr. and MELLOR, A. M. Analytical
calculations for the performance and pollutant emissions
of gas turbine combustors, Comb. Sci. Tech. 4, 101 (1971).
HAMMOND, D. C., Jr. and MELLOR, A. M. Analytical
predictions of emissions from and within an Allison
J-33 combustor, Comb. Sci. Tech. 6, 279 (1973).
HARE, C. T., DIETZMANN, H. E. and SPRINGER, K. J.
Gaseous emissions from a limited sample of military
and commercial aircraft turbine engines, Southwest
Research Inst. AR-816 (1971).
HAWTHORNE, W, R., WEDDELL, D. S,, and HOTTEL,
H. C. Mixing and combustion in turbulent gas jets,
Third Symposium on Combustion and Flame and
Explosion Phenomena, p. 266. Williams & Wilkins,
Baltimore (1949).
HENDERSON, R. E. and BLAZOWSKI, W. S. Aircraft gas
turbine pollution limitations oriented toward minimum
effect on engine performace, Atmospheric Pollution by
Airera/t Engines, AGARD CP No. 125 (1973).
HEYWOOD,J. B. Mass. Inst. Tech., personal communication (1974).
HEYWOOD, J. B. Pollution formation and control in
spark ignition engines, Progress in Energy amt Combustion Science (Chigier, N. A., Ed.), Pergamon, Oxford,
in press (1975).
HEYWOOD, J. B. and MIKUS, T. Parameters controlling
nitric oxide emissions from gas turbine combustors,
Atmospheric Pollution by Aircraft Engines, AGARD CP
No. 125 (1973).
HOTTEL, H. C., WILLIAMS, G. C. and BONNELt., A. H.
Application of stirred reactor theory to the prediction
of combustion chamber performance, Combust. Flame
2, 13 (1958).
HUNTER, S. C., JOHNSON, K. M., MONGIA, H. C. and
WOOD, M. P. Advanced, small high-temperature-rise
combustor program. Vol. I : Analytical model derivation
and combustor-element rig tests (phases 1 and II).
U S A A M R D L Tech. Rep. 74-3A (1974).
JONES, K. H., SAMPSON,R. E. and HOLMES, J. G. The
federal aircraft emission control program: standards
and their basis, J. APCA 24, 23 (1974).
JONES, R. E. and GROBMAN, J. Design and evaluation
of combustors for reducing aircraft engine pollution,
Atmospheric Pollution by Aircraft Engines, AGARD CP
No. 125 (1973).
KOMIYAMA,K., FLAGAN, R. and HEYWOOD, J. B. Mass.
Inst. Teeh., personal communication (1974).
LAPOINTE, C. W. and SCHULTZ, W. L. Measurement of
nitric oxide formation within a multi-fueled turbine
combustor, Emissions fi'om Continuous Combustion
Systems, p. 211. Plenum Press, New York (1972).
LAPOINTE, C. W. and SCHULTZ, W. L. Comparison of
emission indexes within a turbine eombustor operated
on diesel fuel or methane. SAE Paper No. 730669 (1973).

52. LAVOIE,G. A. and SCMLADER,A. F. A scaling study of


NO formation in turbulent diffusion flames of hydrogen
burning in air, Comb. Sci. Tech. 8, 215 (1974).
53. LEEEBVRE,A. H. Design considerations in advanced gas
turbine combustion chambers, Combustion in Advanced
Gas Turbine Systems, p. 3. Pergamon Press, Oxford
(1968).
54. LEFE~VRE, A. H. Factors controlling gas turbine combustion performance at high pressure, Combustion in
Advanced Gas Turbine Systems, p. 211. Pergamon
Press, Oxford (1968).
55. LEFEBVRE,A. H. Pollution control in continuous combustion engines, Fifteenth Symposium (International) on
Combustion, The Combustion Institute, Pittsburgh, in
press (1975).
56. LEFEBVRE, A. H. and FLETCHER, R. S. A preliminary
study on the influence of fuel staging on nitric oxide
emissions from gas turbine combustors, Atmospheric
Pollution by Aircraft Engines, AGARD CP No. 125
(1973).
57. LEMKE, E. E., et al. Air Pollution from Aircraft in Los
Angeles County, Air Pollution Control District, County
of Los Angeles (1965).
58. LIPFERT,F. W. Correlation of gas turbine emissions data.
ASME Paper No. 72-GT-60 (1972).
59. LONGWELL,J. P., FROST, E. E. and WEISS, M. A. Flame
stability in bluff body recirculation zones, Ind. Eng.
Chem. 45, 1629 (1953).
60. Los ANGELES COUNTY AIR POLLUTION CONTROL
DISTRICT. Jet Aircraft Emissions and Air Quality in the
Vicinity of the Los Angeles International Airport (1971).
61. MADOR, R. J. and ROBERTS, R. A pollutant emissions
prediction model for gas turbine combustors, AIAA
Paper No. 74-1113 (1974).
62. MCCREATH, C. G. and CHIGIER, N. A. Liquid spray
burning in the wake of a stabilizer disc, Fourteenth
Symposium (International) on Combustion, p. 1355. The
Combustion Institute, Pittsburgh (1973).
63. MELLOR,A. M. Current kinetic modeling techniques for
continuous flow combustors, Emissions from Continuous
Combustion Systems, p. 23. Plenum Press, New York
(1972).
64. MELLOR, A. M. Simplified physical model of spray
combustion in a gas turbine engine, Comb. Sci. Tech.
8, 101 (1973).
65. MELLOR, A. M. Spray combustion from an air-assist
nozzle, Comb. Sci. Tech. in press (1975).
66. MELLOR,A. M., ANDERSON, R. D., ALTENKIRCH, R. A.
and TUTTLE, J. H. Emissions from and within an Allison
J-33 Combustor, Comb. Sci. Tech. 6, 169 (1972).
67. MERRYMAN, E. L. and LEVY, A. Nitrogen oxide formation in flames: the roles of NO2 and fuel nitrogen,
Fifteenth Symposium (International) on Combustion, The
Combustion Institute, Pittsburgh, in press (1975).
68. MOSIER, S. A., ROBERTS, R. and HENDERSON, R. E.
Development and verification of an analytical model for
predicting emissions from gas turbine engine combustors
during low-power operation, Atmospheric Pollution by
Aircraft Engines, AGARD CP No. 125 (1973).
69. MOS1ER, S. A. and ROBERTS, R. Low-power turbopropulsion combustor exhaust emissions. Vol. II:
Demonstration and total emission analysis and prediction. AFAPL-TR-73-36, Vol. II (1974).
70. MOSIER, S. A. and ROBERTS, R. Low-power turbopropulsion combustor exhaust emissions. Vol. III:
Analysis. AFAPL-TR-73-36, Vol. III (1974).
71. NELSON, A. W. Collection and assessment of aircraft
emissions baseline data-turbine engines. Pratt &
Whitney Aircraft PWA-4339 (1972).
72. NORGREN, C. T. and INGEBO, R. D., Effects of prevaporized fuel on exhaust emissions of an experimental
gas turbine combustor. NASA TMX-68194 (1973).
73. NORSTER, E. R. and LEFEBVRE, A. H. Effects of fuel
injection method on gas turbine combustor emissions,

Gas turbine engine pollution

Emissions from Continuous Combuster Systems, p. 255.


Plenum Press, New York (1972).
74. NORTHERN RESEARCH AND ENGINEERING CORPORATION. Nature and control of aircraft engine exhaust
emissions. PB 187-711 (1968).
75. ODGERS, J. Current theories of combustion within gas
turbine chambers, Fifteenth Symposium (International)
on Combustion, The Combustion Institute, Pittsburgh,
in press (1975).
76. ONUMA, Y. and OGASAWARA, M. Studies on the
structure of a spray combustion flame, Fifteenth
Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, in press (1975).
77. OPDYKE, G. Emissions from Continuous Combustion
Systems, p. 309. Plenum Press, New York (1972).
78. OSGERBY,I. T. Literature review of turbine combustion
modeling and emissions, AIAA J. 12, 743 (1974).
79. PARIKH, P. G., SAWYER, R. F., and LONDON, A. L.
Pollutants from methane fueled gas turbine combustion,
Univ. Cal. Berkeley, Coll. Eng. Rep. No. TS-70-15
(1971).
80. POMPEI, F. and HEYWOOD, J. B. The role of mixing in
burner-generated carbon monoxide and nitric oxide,
Combust. Flame 19, 407 (1972).
81. ROBSON, F. L., GIRAMONT1, A. J., LEWIS, G. P. and
GRUBER, G. Technological and economic feasibility of
advanced power cycles and methods of producing nonpolluting fuels for utility power stations. Rept. prepared
for Nat. Air Poll. Control Admin., U.S. Dept. HEW
(1970).
82. SAWYER,R. F. Atmospheric Pollution by Aircraft Engines
and Fuels. AGARD AR No. 40 (1972).
83. SAWYER,R. F., CERNANSKY,N. P. and OPPENHEIM, A. K.
Factors controlling pollutant emissions from gas turbine
engines, Atmospheric Pollution by Aircraft Engines,
AGARD CP No. 125 (1973).
84. SAWYER, R. F., TEIXEIRA, D. P. and STARKMAN, E. S.
Air pollution characteristics of gas turbine engines,
ASME Trans., J. Eng. Power 91, 290 (1969).
85. SHISLER, R. A., TUTTLE, J. H. and MELLOR, A. M.
Emissions from and within a film-cooled combustor.
Paper No. 74-21. Western States Section/The Combustion Institute (1974).
86. SJ6GREN, A. Soot formation by combustion of an
atomized liquid fuel, Fourteenth Symposium (International) on Combustion, p. 919. The Combustion
Institute, Pittsburgh (1973).
87. SPALDING, D. B. Some Fundamentals of Combustion,
Butterworths, London (1955).
88. SPALDING, D. B. Mathematical models of continuous
combustion, Emissions from Continuous Combustion
Systems, p. 3. Plenum Press, New York (1972).

133

89. STARKMAN,E. S., MIZUTANI, Y., SAWYER, R. F. and


TEIXEIRA, D. P. The role of chemistry in gas turbine
emissions. ASME Paper 70-GT-81 (1970).
90. SWITHENBANK, J., POLL, [., VINCENT, M. W. and
WRIGHT, D. D. Combustion design fundamentals,
Fourteenth Symposium (International) on Combustion,
p. 627. The Combustion Institute, Pittsburgh (1973).
91. TOONE, B. A review of aero engine smoke emission,
Combustion in Advanced Gas Turbine Systems, p. 271.
Pergamon, Oxford (1968).
92. TROTH, D. L., VERDOUW, A. J., WILLIAMS, J. R. and
DODD, R. G. Investigation of aircraft gas turbine combustors having low mass emissions. ASME Paper
No. 74-GT-36 (1974).
93. TUTTLE, J. H., ALTENKIRCH, R. A. and MELLOR, A. M.
Emissions from and within an Allison J-33 combustor.
II. The effect of inlet air temperature, Comb. Sci. Tech.
7, 125 (1973).
94. TUTTLE, J. H., SHISLER, R. A. and MELLOR, A. M.
Nitrogen dioxide formation in gas turbine engines:
measurements and measurement methods, Comb. Sci.
Tech., in press (1975).
95. TUTTLE, J. H., SHISLER, R. A. and MELLOR, A. M.
Purdue Univ., unpublished results (1975).
96. VAUGHT, J. M., JOHNSEN, S. E. J., PARKS, W. M. and
JOHNSON, R. L. Collection and assessment of aircraft
emissions base-line d a t a - - t u r b o p r o p engines (Allison
T56-A- 15). Detroit Diesel Allison EDR-7200 ( 1971 ).
97. VERKAMP, F. J. Detroit Diesel Allison Div., General
Motors Corp., personal communication (1974).
98. VERKAMP,F. J., VERDOUW, A. J. and TOML1NSON,J. G.
Impact of emission regulations on future gas turbine
engine combustors, J. Aircraft 11, 340(1974).
99. VRANOS, A. Turbulent mixing and NOT formation in
gas turbine combustors, Combust. Flame 22, 253 (1974).
100. WADE, W. R. and CORNELIUS, W. Emission characteristics of continuous combustion systems of vehicular
powerplants-gas turbine, steam, Stirling, Emissionsjrom
Continuous Combustion Systems, p. 375. Plenum Press,
New York (1972).
101. WADE, W. R., SHEA, P. I., OWENS, C. W. and MCLEAN,
A. F. Low emissions combustion for the regenerative
gas turbine. Part l : Theoretical and design considerations. ASME Paper No. 73-GT-11 (1973).
102. WAMPLER, F. B., CLARK, D. W. and GAINES, F. A.
Catalytic combustion of C3H8 on Pt-coated monolith,
Paper No. 74-36. Western States Section/The Combustion Institute (1974).
103. ZUKOSI,E. E. and MARBLE,F. E. Experiments concerning the mechanism of flame blowoff from bluff bodies,
Proc. Gas Dynamics Symposium on Aerothermochemistry, p. 205. Northwestern Univ. (1956).

R e c e i v e d 7 J a n u a r y 1975

Das könnte Ihnen auch gefallen