You are on page 1of 12

40th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit AIAA 2004-4192

11 - 14 July 2004, Fort Lauderdale, Florida

Development Needs for Advanced Afterburner Designs

Jeffery A. Lovett*, Torence P. Brogan† and Derk S. Philippona‡


Pratt & Whitney Military Engines, East Hartford, Connecticut, 06108

Barry V. Keil§ and Troy V. Thompson**


Air Force Research Laboratory/Wright-Patterson AFB, Ohio, 45433-7607

The afterburner on a modern aircraft gas turbine engine provides significant thrust
augmentation critical to the performance and mission of tactical aircraft. Higher exhaust
temperatures and survivability requirements for advanced aircraft have resulted in new
constraints on the augmentor design which dramatically changes the design architecture.
Many of the design methods established for afterburners over the past 50 years are
insufficient for the augmentor configurations being developed today. This paper will
describe some of the fundamental technical challenges being faced by augmentor designers,
and will outline critical needs in terms of fundamental combustion sciences and engineering
that needs to be acquired to support new design methods for advanced augmentors.

Nomenclature
f/a = fuel-air ratio
P = pressure
T = temperature
W = flameholder width
V = freestream velocity approaching the flameholder
φ = equivalence ratio

I. Introduction

T he gas turbine engine has transformed the aircraft industry, providing performance levels unachievable using
propeller driven propulsion systems. When maximum thrust is a goal, the thrust augmentation offered by the
afterburner becomes critical to meeting the overall system goals for takeoff, maneuverability and acceleration. The
propulsive efficiency of the afterburner is significantly lower than that afforded by the primary gas turbine operating
on the Brayton cycle, simply because the heat of combustion is released during a lower-pressure part of the cycle.
For this reason, an afterburner is generally not attractive for commercial engines where fuel burn is a dominant
factor. However, for high-speed applications, beyond Mach one, the range in thrust between takeoff and cruise is a
challenge to the propulsion system design and thrust augmentation using an afterburner is generally attractive. A
case in point being the supersonic Concord commercial jet which required an afterburner to meet its system goals.
The development of the gas turbine grew out of the need for pursuit aircraft in World War II that could achieve
higher speed and service ceilings. The J31 was the first production turbojet which was a modified Rolls-Royce
W2B/23C Welland engine, which then became the J33, the first gas turbine engine put into combat on the F-80
Shooting Star. The J47 was developed during the cold-war era and was the first axial-flow turbine and the first to
include an afterburner. Several design iterations of the afterburner were required during and after the engine
development to deal with high altitude operability issues. The J57 was the first dual-spool gas turbine able to
exceed 10,000 lb of thrust used to power the F-100,101 and 102 aircraft. The thrust augmentor enabled both take-

*
Exhaust System Technology Leader, Combustor/Augmentor/Nozzle Module Ctr, MS 163-03, Senior Member.

Staff Aero-Thermal Engineer, Combustor/Augmentor/Nozzle Module Ctr, MS 184-41, Member.

Augmentor Design Chief, Combustor/Augmentor/Nozzle Module Ctr, MS 184-28.
§
Augmentor Technical Leader, AFRL/PRTC, Bldg 490, room 109, Senior Member
**
JSF Program Office, Augmentor/Nozzle CIPT, ASC/LPZ, Bldg 28.

1
American Institute of Aeronautics and Astronautics

Copyright © 2004 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
off and climb acceleration targets. The design systems were becoming reliable enough to enable a successful scale-
up of the J57 to create the J75 engine, which was able to produce over 26,000 lb thrust. The J79 was first delivered
in 1961 for the F-104 and F-4 aircraft, and ignition and dynamics issues with the afterburner lead to the introduction
of a staged afterburner with an advanced pilot system. In addition, high-frequency screech instabilities were
problematical leading to the development of a perforated acoustic-damping liner. The need for increased range
drove the development of augmented turbofan engines which introduced new challenges for the afterburner design
since a cool fan stream and hot core stream now actively mixed at the inlet to the augmentor. Difficulties with
translating the turbojet design systems were ultimately overcome in developing the TF30 that was used on several
tactical aircraft. The TF30 afterburner was the predecessor to the F-100 engine delivered in 1975 to power the F-15
Eagle high-performance interceptor. Total thrust was similar to the J75 but the thrust-to-weight ratio had grown to
over seven. Again, significant development effort was needed over two decades to manage flame stability and
combustion dynamics, highlighting the lack of robust design rules for the afterburner.
The brief history offered above illustrates how past augmentor development occurred primarily after the engine
was delivered for service, and with each significant change in engine architecture. Today, high exhaust
temperatures and new survivability requirements are generating a major revolution in afterburner design which will
require another substantial investment in technology development to ensure reliable and robust designs to meet ever
increasing system goals. The escalating costs of engine testing and full-scale development has focused attention on
upfront development to avoid iterations late in the engine program. For the augmentor, this equates to a need for
much better physical understanding of the complex combustion processes involved and the development of robust
design rules and tools to guide the design architecture early in the development process. The purpose of this paper is
to outline the fundamental combustion science challenges to meet these objectives.

II. Description of the Typical Afterburner


The overall function of the afterburner is relatively simple. It is an engine component which simply acts to
consume the remaining oxygen in the turbine exhaust and generate usable thrust by increasing the exhaust
temperature and velocity. Because the inlet stream to the augmentor is vitiated air, i.e., partially consumed oxygen,
the conditions for combustion are significantly different than those considered for the main combustion system.
Table 1 lists a comparison of typical inlet conditions for the main combustor and the afterburner in a gas turbine
engine. From the point of view of combustion, only the operating pressure represents a lower capability for the
afterburner. The flow velocity and temperature are considerably higher for the afterburner, affecting directly the
fundamental fuel injection and combustion processes associated with the system. Furthermore, the oxygen content
present in the afterburner inlet stream is
significantly reduced relative to the main
combustor, resulting in reduced chemical reaction Table 1. Approximate Typical Inlet Conditions for
rates. These fundamental differences will be a Gas Turbine Combustor and Afterburner.
discussed in detail in the following sections of this
paper. Suffice it to say, that the continual growth in Inlet Condition Main Combustor Augmentor
thrust performance of the gas turbine engine Temperature (°C) 350 - 650 650 - 1050
manifests itself primarily in terms of increased inlet Pressure (atm) 10 – 30 0.5 - 6
temperature and reduced oxygen content for the Velocity (m/s) 30 - 60 150 - 250
augmentor which profoundly effects the design Oxygen (% vol) 21 12 – 17
architecture of the system.
The other aspect which is having a profound impact on the design architecture of the afterburner is the
survivability requirements for modern tactical aircraft. Since the augmentor cavity represents a primary direct
source of observability for the aircraft, substantial requirements are placed on the augmentor system. To meet these
system objectives requires a radically new design architecture for the system. In order to confidently accommodate
these new requirements, improved fundamental data and analytical tools are needed.
The relatively simple function and purpose for the augmentor resulted in a fairly simple design architecture in
past designs. Past designs strived for generating good combustion uniformity & efficiency while minimizing
weight, cost & dry pressure loss. Referring to Figure 1, fuel is typically introduced into the augmentor using a series
of fuel manifolds rings or radial struts with a large number of fuel injection sites. This arrangement offers excellent
fuel distribution uniformity, or premixing, which in turn provides excellent combustion efficiency. The flame is
typically stabilized using an array of bluffbody flameholders which are made of V-shaped gutters providing robust
fluid recirculation zones in the flow to anchor the flame in space within the augmentor cavity. The array of
flameholders is arranged to provide combustion uniformity while minimizing the effects of flow blockage and

2
American Institute of Aeronautics and Astronautics
pressure drop so as to introduce a minimum pressure loss in “dry” or afterburning mode. This is important since
flow pressure loss from the afterburner system directly decreases the thrust output in non-afterburning mode where
cruise operation and fuel consumption is
most important. Typically, a pilot ring is FUEL FUEL INJECTORS AUGMENTOR LINER
FAN AIR`
used to ignite and anchor the flame on the
flameholders and provide robust operability
CORE AIR
over the range of conditions required. The FLAMEHOLDERS
pilot flame is often directly fueled and
PILOT
arranged to be very stable so that it anchors LP TURBINE
or carries the rest of the combustion process
over the range of conditions for which the
augmentor must operate. The flameholders Figure 1. Depiction of a typical afterburner configuration
are responsible for spreading the flame
for modern in-service aircraft engines.
uniformly across the duct to provide
sufficient combustion efficiency.
The design challenges associated with past augmentor designs usually revolved around maximizing the flow and
combustion uniformity and providing good operability, reliability, and durability. The durability issues were
typically associated directly with the light-weight, thin, sheet metal construction methods operating at very high
temperatures and convective heat transfer rates. Proper design methods and techniques have been developed to
accommodate the large thermal growth and thermal stresses occurring in afterburner structures. The reliability of
the system is essentially associated with providing a dependable ignition process and staging the introduction of fuel
so that the transient pressure rise experienced when the afterburner fuel first ignites does not create stability or
operational problems in other parts of the engine system. Fuel injection and combustion uniformity was typically
developed empirically through parametric testing and involved the trade-off between performance and durability.
Operability for afterburners generally refers to both static and dynamic stability. Static stability is associated
with the ability of the system to hold flame steadily on the flameholders over the entire range of operating conditions
related to the aircraft flight envelope. Dynamic stability speaks to the unsteady character of the flame where strong
oscillations or instability of the flame can create problems with high cycle fatigue or high local heat transfer
enhancement. Operability has always been a challenge to augmentor designers (Bonnell, et al., 1971) because the
afterburner combustion system must operate over a substantial range of conditions to satisfy the aircraft flight
envelope. Figure 2 signifies an aircraft flight envelope, defining the range of conditions that the afterburner system
must accommodate. On the x-axis is the aircraft flight Mach number which affects the inlet conditions (temperature
and pressure) into the engine, and therefore, into the augmentor. On the right side of the envelope, the flight speed
is highest and the operating pressures and temperatures reach maximum levels. In the upper left-hand corner of the
envelope, corresponding to the highest flight altitudes, the lowest operating pressures and temperatures are generally
found. The contours shown in Figure 2 illustrate
that typically regions of constant operating pressure
extend diagonally across the flight envelope
ULHC
corresponding to the ambient pressure and ram
pressure behavior. The range of operating pressure
and fuel flowrate for the afterburner from upper-left
to the lower-right extents of the flight envelope is
typically on the order of 10:1. This large range in ALTITUDE INCREASING
operating pressure represents the most fundamental CHAMBER
PRESSURE
challenge for combustion engineers designing
augmentors since the fuel injection and combustion
processes are strongly dependent on pressure. As
engine thrust capability increases, core vitiation
level increases result in reduced partial pressures
FLIGHT MACH NUMBER
exasperating low-pressure combustion kinetic
behavior.
Figure 2. Depiction of a typical aircraft engine flight
envelope and contours of constant pressure.

3
American Institute of Aeronautics and Astronautics
III. Challenges for Advanced Augmentor Designs
The ratio of thrust to weight for the gas turbine engine has increased substantially over the past 50 years,
enabling impressive growth in aircraft performance. This increased performance is largely the result of increases in
the turbine inlet temperature afforded by turbine material and cooling improvements and increased main combustor
operating temperatures. This equates to increased turbine exhaust temperature and decreased oxygen concentration
entering the afterburner. The injection of fuel into the
10
turbine exhaust stream to power the augmentor is limited O=1
7 atms
by autoignition of the fuel spray which defines a
fundamental limitation regarding the distance that can be IGNITION
IGNITION
DELAY RANGE
used to premix the fuel prior to the flameholders (ref.
IGNITION
Figure 1). Figure 3 illustrates the approximate ignition DELAY TIME
delay time expected for jet fuels as a function of the gas (msec) 1
temperature derived from studies by Colket and
RANGE OF FUEL
Spadaccini (2001, 1971). Peak turbine exit temperatures TRANSPORT TIME
including radial profile can exceed 1300K in high
performance engines. For these high temperatures, NO
IGNITION PEAK
autoignition times can be less than one millisecond. TURBINE
Estimated fuel spray transport times corresponding to EXIT TEMP
0.1
fuel injector separations greater than 50mm are also of
800 1000 1200 1400 1600
the order of 1 msec. Thus, the short autoignition time
results in maximum premixing distances on the order of GAS TEMPERATURE (K)
50 mm which significantly limits the ability to generate
a uniform distribution of fuel at the flameholder and Figure 3. Ignition delay time as a function of
over the large cross-sectional area of the augmentor. mixture temperature.
The higher inlet temperatures in today’s afterburners also challenge the temperature limits of metal alloys. This
directly affects the design of the flameholders and fuel injectors, which are immersed in the turbine exit flow. To
accommodate this requires active cooling of the flameholder and fuel injector using air extracted from the fan stream
to keep the metal temperatures at an acceptable level for long-term durability. This substantially complicates the
design and construction of the flameholder and fuel injector components, requiring multiple-wall construction and
increased component weight. This fundamental design constraint has lead to the integration of the fuel injectors
with the flameholder system (Wadia, et al., 2001), as shown in Figure 4. Furthermore, cooling of the turbine support
struts, combustion liners and exhaust nozzle are also
similarly challenged, requiring a direct supply of
cooling air from the fan stream to actively cool all FUEL FUEL INJECTORS CROSS-SECTIONAL VIEW
the hot gas-path parts. To accommodate the
FUEL INJECTOR
various active coolings in the exhaust system FAN AIR

requires a good controlled supply of cooling air


provided by directly attaching the augmentor liner CORE AIR FLAMEHOLDER FLAME

to the turbine OD wall. This is often termed a “tied-


liner” system since the hot-path liners are tied 50 mm
together to create a separate flow path for the fan air
stream. Although this is largely beneficial for Figure 4. Depiction of a close-coupled fuel injector and
cooling control, it creates a stiff coupling between flameholder system.
the augmentor and the engine core streams.

A. Fuel Delivery System Requirements


The fuel injection system in most afterburner designs has been comprised in the past of a relatively simple
system of tube manifolds since the low core temperature allowed unclooled fuel injectors to be used. In the close-
coupled architecture required for advanced augmentors, the fuel manifold is placed within the air-cooled
flameholder and the fuel spray is typically injected normal to the surface and also normal to the flow. The
requirement to keep the fuel transport time below the autoignition limit discussed above dictates the maximum
allowable upstream distance for the fuel injector. The premixing length is substantially reduced and the fuel spray
directly interacts with the flame being stabilized on the flameholder. Therefore, the fuel spray characteristics can
have a much larger influence on the flame behavior in the close-coupled system. For example, the fuel jet
penetration will vary at different operating conditions which will directly affect the local fueling of the flameholder.

4
American Institute of Aeronautics and Astronautics
Figure 5 shows the range of fuel jet penetration calculated for a nominal fuel orifice over a flight envelope. From
this example, it can be seen that the fuel jet penetration, and therefore the relationship of the fuel jet to the
flameholder, can vary by over an order of magnitude across the flight envelope. This type of dependency had much
less influence in past designs since there was more mixing time to even the fuel distribution out before reaching the
flameholder. With the close-coupled arrangement, it is critical to understand the details of the fuel spray as it relates
to the flame zone on the flameholders, and how it changes through the flight envelope.
A straightforward method of achieving a uniform
fuel distribution is to use many fuel injection sites to
10 mm
cover the flow cross-sectional area. With the fuel 15 mm
injector integrated with the flameholder, weight and flow 20 mm

blockage become coupled with the number of fuel 25 mm


injection sites used. Furthermore, the requirement to
ALTITUDE
duct cool fan air to the injector/flameholder system LIQUID JET
PENETRATION
30 mm

makes a radial flameholder system most practical. To


increase fuel injection sites is then equivalent to 35 mm
increasing the number of flameholders and the
associated weight and blockage. Together, these design
constraints indicate that methods to improve fuel jet FLIGHT MACH NUMBER
penetration and mixing may be required to optimize
combustion performance and system weight using a Figure 5. Example of range in fuel jet penetration
minimum number of fuel injection sites and expected for a simple-orifice fuel injector.
flameholders.

B. Flame Stabilization System Requirements


In current afterburner designs, flame stabilization is achieved by providing bluffbody recirculation zones behind
a system of flameholders to ignite the oncoming fuel-air mixture. The static stability or blowoff limit of a flame
stabilized behind a bluffbody depends upon the wake properties and generally follows a relationship of the form
[Ozawa, et al., 1970]

 V  1 
  1.5 
= f (φ )
 W  PT 

where V the approach velocity, W is the flameholder width, P the pressure, T the approach temperature and φ the
fuel-air equivalence ratio. This expression is related to the ratio of the fluid mixing time to the chemical reaction
time. The primary factors influencing the stability are then the fuel-oxidant ratio of the mixture, the flow speed, and
the size of the flameholder. In general, we have said that the turbine exit temperature has increased in advanced
engines, so clearly this is beneficial for overall stability of the flameholder system. But the higher core temperature
is essentially the result of operating at a high fuel-air ratio in the main combustor, and so the oxygen concentration
in the turbine exit is reduced considerably. This higher vitiation level acts to reduce the chemical kinetics and so
counteracts the benefit of the higher gas temperature. Chemical kinetics calculations of the fundamental reaction
time for conditions relevant to an advanced engine cycle relative to in-service engines indicates that the net effect of
higher turbine operating temperatures decreases kinetic times and therefore benefits combustion stability in the
augmentor.
Although the increases in turbine temperature on their own would improve combustion rates, other aspects of the
new augmentor architecture can dramatically change the mixture present at the flameholders and therefore strongly
impact stability. Foremost is the fact that the mixture cannot be assumed to be well premixed because the discrete
fuel injection sites are closely coupled to the flame as discussed above. Instead, fuel from the individual fuel jets
directly feeds the flameholding region and control of the mixture in the near-field flameholder region is much more
difficult. In past augmentor designs, the results of premixed flameholder experiments (ref. Ozawa, 1970) could be
used to define the stability limits as a function of bulk operating conditions and flameholder geometry. Since the
fuel mixture in the augmentor was relatively well premixed, these experimental correlations very successfully
characterized the basic stability limits. These data are no longer so valuable for systems where the fuel mixture is
non-uniform relative to the flameholder and core-stream vitiation levels are significant.

5
American Institute of Aeronautics and Astronautics
The non-premixed nature of the flow in
CORE
advanced augmentors makes it necessary to VITIATION
understand the local properties of the flameholder
FUEL SPRAY FLAMEHOLDER COMBUSTION
wake and shear-layers. The reaction zone attached ENTRAINMENT WAKE COMPOSITION RATE
to the flameholders is generally associated with the
fluid shear-layers between the core stream and the WALL
COOLING AIR
flameholder wake. Since the bluffbody wake PILOT
typically generates very high mixing rates, the STREAM
properties in the flameholder wake are considered
equivalent here (Williams, 1955). Several streams Figure 6. Different streams which contribute to the
can contribute to the local wake properties as flameholder wake composition.
illustrated in Figure 6. In addition to the core
stream and the discrete fuel spray from the
augmentor fuel injectors, the cooling air used to cool the walls of the flameholder structure can also directly impact
the flameholder wake mixture. Thus, the mixture composition which determines the combustion stability of
flameholder is now a complex mixture from several sources with varying temperature and oxygen content. The
temperature and fuel and oxygen concentrations which are vitally important in governing the combustion process
must be determined accurately to predict the stability performance of the system.
Once again, the entrainment of cooling air into
the flameholder wake involves competing effects.
Since the air is derived from the fan stream, this air
will offer added oxygen to the vitiated mixture
which can increase combustion reaction rates. 300K COOLING AIR
However, the fan stream is relatively cool air which
can reduce the chemical kinetic rates if the mixture CHANGE IN
temperature is depressed. Figure 7 shows examples KINETIC TIME 500K COOLING AIR
of the change in combustion time calculated when (10-6sec)
air at various temperatures is mixed with a vitiated
core stream to create a stoichiometric fuel-air 700K COOLING AIR

mixture. These simulations show that there is a


trade-off between the effects of temperature and
fresh oxygen as it impacts the combustion kinetics % AIR ADDED
time. Modern engines typically operate at higher
pressure ratios, and combined with high ram Figure 7. Estimate of kinetic time for a mixture of fuel
pressures, results in significant increases in the fan and air with an added stream of cooling air.
air temperature. A good understanding of the wake
conditions will be needed to ascertain the effects of
these features on the system static stability.

C. Pilot & Ignition System Requirements


A pilot system is typically used to ignite and anchor the flame on the flameholders and provide robust operability
over the range of conditions required. The most common pilot system is a form of bluffbody flameholder itself,
arranged in an annular ring, and contacting the main flameholders in a way to transfer the pilot flame products to the
flameholders to enhance the combustion process. The pilot flameholder will typically have its own fuel delivery
sources and a spark igniter. A favorable design trade is usually received by minimizing the main flameholder
blockage to reduce non-afterburning pressure loss as it directly improves specific fuel consumption. The pilot is
then crucial to enhance the stability of the flameholders. Given the aforementioned difficulty in designing a close-
coupled flameholder system with robust static stability over the full range of conditions, it is likely that the pilot
system will play an even more crucial role in the design of new augmentor systems. The pilot system will be subject
to many of the same constraints described for the flameholder system, most notably that the tied-liner aspect of the
augmentor configuration will mean that the pilot itself will be subject to the strong pressure and flow variations
which can exist over the operating envelope and during engine transients. The composition of the pilot is another
stream which will contribute to the local flameholder wake conditions and therefore to the flameholder stability.
Effective integration of the pilot gas with the basic flameholder combustion conditions will be crucial to the design
of advanced augmentors.

6
American Institute of Aeronautics and Astronautics
D. Dynamic Stability Requirements
Combustion instabilities, in the form of screech (high frequency) and rumble (low frequency), have always
impacted the design of thrust augmentors. The theory of combustion instability relates the gain in the coupled
thermo-acoustic system to the overall heat release occurring in the combustion chamber. The system pressure
continues to rise in advanced engine designs, which means higher energy density in the augmentor and therefore a
higher propensity for driving combustion instabilities. The driving must be balanced by damping in the system.
Some of the damping is associated with paths for acoustic energy to escape the combustion chamber through
openings and coupled volumes. Much of the
damping opportunities are eliminated with the
attached or tied liner architecture, making the 0.8 m
Diameter
combustion chamber now very simple and rigid. Chamber 50 mm Backing Height
Additional damping has always been required in
afterburners and is typically provided by a
perforated liner which derives its damping influence AMPING
through a Helmholtz resonator response. The
damping response of perforated liners typically
increases with increasing frequency, as illustrated in 0.6 m
Diameter
Figure 8. As advanced engines increase in size with Chamber
core growth, the fundamental acoustic frequencies
0 1kHz 2kHz
decrease in proportion. This makes it more difficult
to get adequate damping at the acoustic frequency FREQUENCY
of the chamber. Thus, the risk for combustion
Figure 8. Illustration of the damping response of a
instabilities is significantly higher for advanced
perforated augmentor liner.
engines having less overall damping and higher
energy release rates.
One of the fundamental difficulties with advancing the understanding of dynamic stability in afterburners is
related to the fact that laboratory rig tests can not faithfully replicate the physics of the instabilities because of the
close dependency on the acoustic environment. The frequencies and mode characteristics are directly related to the
presence of a large cylindrical or annular combustion chamber with a diameter on the order of a meter.
Furthermore, the upstream acoustic characteristics in the engine is dependent on the complex impedance associated
with the engine turbomachinery. Laboratory rigs are typically limited to less than one-third this size, and higher
frequencies achievable using sub-scale tests are subject to different coupling and damping mechanisms compared
with the engine environment.

IV. Research & Development Needs


The fundamental changes in the basic architecture of advanced afterburners described above represents a
revolutionary change in the design. Many of the design rules and guidelines for which augmentor designs were
based are no longer appropriate for the new systems, which are very highly integrated and challenge the premixed
combustion assumptions adopted in the past. An acceleration of research activity is needed in this area to establish
the fundamental database and design guidelines for the close-coupled augmentor architecture with high core
temperature and vitiation conditions. This need is amplified by the substantial rise in development costs for full-
scale engine tests, highlighting the need for reliable design rules and tools to avoid costly iterations at full scale.
Some of the research & development needs to support the development of new augmentor systems are described
next.
The literature database for the close-coupled flameholder and close-coupled fuel injection was largely an
outgrowth of the turboramjet ramburner development program, initiated by the US airforce research labs in 1987
(Wotel and Gallagher 1989). General combustion characteristics and fuel injection were studied by Hautman et al.
(1990a, 1990b), and Wotel et al. (1991). The complete non-reacting flowfield was documented using three-
component laser velocimetry with both mean and turbulence quantities for a good benchmark database (Raffoul et
al. 1995). Mixing characteristics under non-reacting and reacting conditions were measured by Gould et al. (1997)
using PLIF techniques. Shahnam et al. (1998) later studied the combustion instability cycle of the same
configuration that arose from an interaction of the fuel jet being entrained with vortex shedding and unsteady
combustion in the wake. These works have illustrated a number of the issues for the close-coupled fuel injector and
flameholder design at a laboratory environment; all were conducted under ambient inlet conditions, with the

7
American Institute of Aeronautics and Astronautics
exception of Wotel et al. (1991). New work is required in the areas of fuel spray, flame stability and non-uniform
fueling, shear-layer dynamics, flame propagation, design tools, and full-scale engine diagnostics.

Fuel Spray Behavior at High Temperature & Ma# Conditions


Fundamental measurements of the fuel spray characteristics associated with high gas temperatures and flow
Mach numbers are needed. Because of the low cost and simplicity of the simple jet in cross flow, this has been the
method of choice for augmentors with various injection angles into the cross-stream to affect penetration and
mixing. There have been substantial studies on two-phase liquid jets, many of which were motivated by research
activity on turbo-ramjet systems (Hojnacki, 1972; Wotel et al., 1991). However, very little data exists for the
conditions relevant to afterburners, where strong evaporation can affect jet penetration and droplet sizes (Hewitt &
Schetz, 1983; Inamura et al., 2001). In fact, many studies concentrated only on the macroscale penetration aspect of
the liquid jet for normal jets, reviewed by (Wu et al., 1997) and angled jets, reviewed by Fuller et al. (2000). Only a
few studies have focused on spray structure such as droplet size, mass flux, and velocity, or more complete far-field
droplet statistics. Since the spatial distribution of fuel is needed to fully determine local fuel entrainment into the
flame zone, the full evolution of the three-dimensional fuel spray must be characterized in the near-field at 20 to 100
injector diameters from the injection location (Leong & Hautman, 2002).
Measurement data for the simple jet in cross-
flow at elevated temperature is surprisingly sparse.
5
Figure 9 shows the disparity of the conditions
Augmentor
relative to where augmentors must operate.
Normalized Temperature (T / 530R)
Conditions
Existing data cover appropriate pressures since
4
afterburner operating pressures are not as high as
other combustors. But augmentor temperatures are ρ /ρ STP = 1
well above conditions for the highest temperature
3
data which currently exists. Hewitt & Schetz (1983)
simulated the temperature effect by using Freon
injectant, and Inamura et al. (2001) measured only
2
penetration values at elevated temperature from (Hewitt & Schetz, 1983)
Jet-fuels Saturation
600K to 900K. The latter showed penetration
differences of more than twice that for ambient
Turbo-Ramjet
conditions at 50 jet diameters downstream and with 1
(Wotel et al., 1990)
the same momentum flux ratio. It is clear that
measurements of the fuel spray at appropriate
conditions is one of the major missing elements to 0
0.1 1 10 100
develop precise afterburner engineering methods.
Pressure (atm)
With the complexity of evaporation included in the
fuel distribution and concentration field, further
Figure 9. Range of fuel spray conditions relevant to
progress requires a combination of both modeling
augmentor conditions compared to existing data which
and experimental work to identify liquid and vapor
includes fuel spray distribution information.
fractions in the near-field. These types of data are
difficult to acquire, however, the validation of such
processes will enable the extension from the laboratory environment to the engine environment in complex three-
dimensional flow fields. Applications of PDPA and LV have been successful to quantify liquid phase properties
even at reacting conditions (Locke et al. 1998) and recent advances with PLIF and PIV methods may also prove
useful for non-reacting full-field measurements. Experiments must distinguish the proper combination of Weber
number (aerodynamic forces to surface tension) and Ohnesorge number (viscous forces to surface tension) to
establish the proper droplet formation processes, but also match the transfer number characteristics for droplet
evaporation.
For simple pressure-atomized liquid jets, the fuel spray penetration and distributions change dramatically over
the flight envelope. Since the flame zone is directly affected by the fuel distribution in the close-coupled system,
advanced fuel injection methods which are less sensitive to injector pressure drop could offer significant advantages.
Another approach is to directly fuel the wake recirculation zone that avoids the dependence on penetration. Early
work has been done with injection methods in v-gutter flameholders (Mellor, 1980), but local details are lacking to
capitalize on the proper integration of aerodynamics and fuel concentration to be used in advanced designs. Fuel
injector durability and coking are additional practical issues with this approach.

8
American Institute of Aeronautics and Astronautics
Flame Stability as Influenced by Local Fuel Distribution
The close-coupled nature of the fuel injection results in a non-uniform distribution of fuel at the flame zone.
Studies to characterize the influence of fuel distribution along and normal to a bluffbody flameholder are needed to
determine the sensitivity of stability to fuel injector spacing and spray distribution. This is a multi-variable problem
since the fuel spray involves different vapor fractions and spray droplet size distributions as well. The lateral spacing
of discrete injectors creates an interrupted flame extending from the flameholder that at some point behaves as
individual jets bounded by quenching regions. It is reasonable to expect that there is a different near-field behavior
progressing from the continuous flame sheet to isolated burning jets. A well conceived study to determine how
important these very local characteristics are to flame stability would be extremely useful in defining practical limits
for the fuel delivery system.

Shear-Layer Combustion Characteristics


The actual flame zone is closely associated with the fluid shear layers surrounding the flameholder recirculation
zone (Williams, 1955). The characteristics of a reacting shear layer are still not well understood, particularly for the
case of a vitiated core stream which reduces the overall heat of reaction and temperature rise. The resulting density
ratio between the free-stream and reacted wake is 2.5 to 3 in the augmentor case, compared with a value of about 7.5
for experiments that are commonly done at ambient conditions (300K, Hydrocarbon fuels). This significantly
changes the dilatation effects and to a lesser extent baroclinic torque in the reacting shear layers, which in turn
affects the fundamental behavior of the large-scale unsteadiness. Studies to quantify the flame structure in reacting
shear layers for the vitiated, high-velocity conditions relevant to afterburners will be extremely useful for validating
computational fluid dynamic methods to simulate the flame zone. Law (1992) describes areas of limited
understanding in high-speed flameholding that are still relevant today. In addition, the unsteady character of the
shear layer, which involves the damping effects of heat release (Mehta and Sorteriou, 2003), is not well understood
but is critical to determining the mixing and entrainment of fuel and air streams governing flame stability and
dynamics.
Studies must clearly delineate consideration for the various turbulent combustion regimes. For example, at sub-
atmospheric conditions flames are kinetically limited and their length scale can be large enough for the flamelet
regime and potentially wrinkled flames to be the appropriate paradigm. At ram conditions, pressures are high
enough where the combustion is mixing limited and auto-ignition behavior can be important. Studies to identify the
fundamental combustion processes for afterburner flames is needed to frame the combustion modeling assumptions.
Knowledge of the basic combustion processes will form the basis of reduced-order models that can be used in order
to make the three-dimensional solution tractable for engineering design.

Flame Spreading as Influenced by Static Stability, Vitiation and Turbulence Scale


The combustion efficiency associated with the augmentor is very important to the maximum thrust of the
propulsion system. Global flame spreading correlations from the past provided a good guideline for estimating the
augmentor efficiency and defining the required system length. However, the ability to more accurately predict
efficiency and understand the variation in performance across the flight envelope will require more complete
understanding of flame spreading and propagation. The basic influence of turbulence characteristics and air vitiation
on flame propagation is needed to more accurately predict flame spreading in an augmentor environment. In
addition, it is generally observed that the combustion efficiency is impacted by the static stability of the
flameholders since the initial conditions of the flame are directly linked to the state of the near-field flame zone.
Studies to better understand this impact will link the tools for achieving stability with the system performance
predictions. Furthermore, the overall evolution of the heat release down the augmentor duct plays a central role in
models for combustion instability.

Design Methodologies and Tools


The fluid dynamic and combustion processes involved with the afterburner are very complex and in themselves
strongly coupled and integrated. As the design architecture becomes more compact with integrated fuel injectors,
active air-cooling at the flowpath surfaces, and reliance on piloting behavior, the ability to use traditional design
rules becomes weaker or simply inaccurate. The computational fluid dynamic methods available today cannot
accurately simulate the complex processes associated with two-phase liquid-fuel injection, ignition, flame stability
and combustion dynamics. Innovative design methodologies are needed which leverage available analytical methods
and understanding to create tools which adequately define performance as functions of the various system variables.
Such system-level tools applied over the range of conditions required of the afterburner can be used to define
optimum configurations offering good system performance and robustness. To facilitate such evaluations, it is likely

9
American Institute of Aeronautics and Astronautics
that such tools will be “reduced-order” in some element of the physics for application across a range of operating
conditions. Such design methodologies/tools for ignition need to account for the impact of local velocity and
mixing since flame extinction can occur at the high flow velocities in augmentors. An adept design methodology for
static stability is needed which captures the impacts of the various local characteristics related to the fuel injection
and cooling flows as they interact with the flameholder system. Tools to accurately simulate flame spreading to
predict combustion efficiency and heat release distribution should be amenable to CFD methods once the impact of
turbulence, vitiation and initial flame composition are quantified and made available to scale results. Acquiring a
design methodology to deal with dynamic stability is perhaps the most difficult challenge at the present time because
of the lack of a unified mechanism to describe thermo-acoustic coupling.
In future applications, significant progress in CFD is anticipated to enable further insights to the controlling
physics for augmentor systems in the development stages of design. Recent examples have shown moderate success
in capturing blowout phenomena, which include unsteady RANS with detailed PDF chemistry models (Frolov et al.,
2000a, 2000b), LES with a simple equilibrium/extinction combustion model (Kim et al., 2004), and LES with a
more general turbulent-combustion model (Menon, 2003). Black and Smith (2003) showed promising transient
blowout simulations for a liquid fueled combustor using both unsteady RANS and LES with a single step reaction
and two variable PDF method for turbulent-chemistry interaction. From these studies, it is clear that unsteady CFD
methods show promise for predicting complex blowout behavior.
It is generally accepted that the flame blowout process is strongly linked to the flow unsteadiness. LES methods
show promise to capture the large-scale unsteady structures in the flow field which can provide accurate unsteady
statistics and time averaged quantities critical to capturing the transport processes important to blowout. These tools
are becoming mature enough to consider for engine design, but they require large computational resources and
further validation is needed to establish guidelines for resolution versus accuracy. Detailed kinetics calculations
coupled with LES are probably not tractable in the near term, so further progress is needed to develop
computationally efficient models that can adequately accommodate the wide range of turbulent combustion regimes
that can co-exist in full-scale augmentors. The presence of fuel spray exacerbates the range of combustion regimes
and can even introduce broken flame zones controlled by a non-uniform fuel/air mixture. Development has started
to provide spray sub-models to simulate the distributed liquid jet break-up and transport processes for CFD
applications (Madabhushi et al., 2004) but their application to elevated temperatures is currently paced by available
data as discussed above.
Applications of CFD to combustion dynamics has also focused heavily on the near-field wake behavior as the
root of the acoustic-coupling mechanism, which again will likely require finite-rate chemistry and time-accurate
(resolved) unsteadiness. However, combustion dynamics simulations using CFD also requires proper treatment of
the acoustic field (Menon, 2003), acoustic boundary conditions, and models to properly treat the impact of acoustics
on the fuel spray (Sujith et al., 1999). Limited work has been done at full-scale problems due to the difficulty and
expense of the calculations. Again, reduced-order modeling coupled to CFD is a path to obtain useful tools since
computing capacity is likely to be several years away.

Engine Measurements and Diagnostics


Detailed diagnostic measurements in a full-scale afterburner are needed to obtain a better description of the fluid
and combustion behavior and to validate capabilities to simulate the three-dimensional flow fields in the engine. The
harsh condition of an afterburner has precluded any local measurements within the combustion system to date.
However, some benchmark measurements will be critical to validate physics-based tools to guide augmentor design.
Innovative instrumentation approaches and possibly use of optical techniques need to be applied to the augmentor to
acquire this important information. The combustion behavior in the close-coupled afterburner is clearly dependent
on very local characteristics; therefore, engine behavior and performance will be strongly coupled to the three-
dimensional flow field in the engine. Engine validation data will be key to developing reliable methods to account
for the true engine environment.
Moreover, both the research and design communities need a consistent picture of the controlling physics of the
real design environment. The rig test and model simplifications will continue to be a fact of component
development in afterburners, as such, the engine diagnostics described above will also help to delineate which
approximations are and are not proper to faithfully replicate the controlling physics to the processes of efficiency,
static stability, and combustion dynamics.

10
American Institute of Aeronautics and Astronautics
V. Summary
The increased exhaust temperatures and survivability requirements for advanced augmentors has initiated a
revolutionary change in the design architecture of gas turbine afterburners. The new constraints require
tremendously integrated systems which closely couple the function of the afterburner with the rest of the engine
system. The function and operability of the augmentor is governed by very local, complex aerodynamic processes
which must be better understood to enable reliable engine development and to avoid costly design iterations on the
engine late in the engine program. A description of the fundamental changes in configuration and the research and
development needs to support this new approach have been discussed.

Acknowledgments
The authors would like to acknowledge the collaborations of Dr. Donald Hautman and Dr. Meredith Colket, III
from the United Technologies Research Center for their contributions to this topic.

References
Black, D. L., Smith, C. E., (2003), “Transient Lean Blowout Modeling of an Aero Low Emission Fuel Injector,” AIAA Paper
2003-4520, Joint Propulsion Conference, July 2003.
Bonnell, J. M., Marshall, R. L., and Reicke, G. T., “Combustion Instability in Turbojet and Turbofan Augmentors,”
AIAA/SAE 7th Propulsion Joint Specialist Conference, Salt Lake City, Utah, AIAA Paper No. 71-698, 1971.
Colket, M. B, III, and Spadaccini, L. J., “Scramjet Fuels Autoignition Study,” Journal of Propulsion and Power, Vol. 17, No.
2, 2001, pp. 315-323.
Frolov, S.M., Basevich, V.Ya., Belyaev, A.A. (2000a), “Mechanism of Turbulent Flame Stabilization on a Bluff Body,”
Chemistry and Physics Reports, Vol 18, No. 8, pp. 1495-1516.
Frolov, S.M., Basevich, V.Ya., Belyaev, A.A. (2000b), “Modeling of Turbulent Flame Stabilization on Bluff Bodies,”
Chemistry and Physics Reports, Vol 18, No. 8, pp. 1683-1704.
Fuller, R.P., Wu, Pei-Kuan, Kirkindall, K.A., (2000), “Effects of Injection Angle on Atomization of Liquid Jets in Transverse
Airflow,” AIAA Journal, Vol 38. no. 1, pp. 64-72
Gould, Richard D. (1997) “Three-Dimensional Mixing Study of Reacting and Isothermal Flow Behind a Bluff Body
Flameholder With Normal Fuel Jet Injection,” ASME Paper FEDSM97-3097, 1997 ASME Fluids Engineering Division Summer
Meeting, June, 1997.
Hautman, D.J., Pfau, D.J., Anderson, T.J. (1990a), “Combustion Tests in a Ramjet Test Combustor,” 27th JANNAF
Combustion Symposium, Vol II, pp. 421-434, Cheyenne, WY.
Hautman, Donald J., Rosfjord, Thomas (1990b), “Transverse Liquid Injection Studies,” AIAA 90-1965.
Hewitt, P.W. and Schetz, J.A., (1983), “Transverse Jet Break-up and Atomization with Rapid Vaporization along the
Trajectory,” AIAA 83-1400, January 1983
Hojnacki, J.T. (1972), “Ramjet Engine Fuel Injection Studies,” Aero Propulsion Lab., AFAPL-TR-72-76, Wright-Patterson
AFB, OH, 1972.
Inamura, T., Takahashi, M., Kumakawa, A. (2001), “Combustion Characteristics of a Liquid-Fueled Ramjet Combustor,” J.
of Propulsion and Power, Vol. 17, No. 4, pp. 860-868.
Kim, Won-Wook, Lienau, Jeffrey J., Van Slooten, Paul R., Colket, Meredith B. III, Malecki, Robert E., Syed, Saadat, (2004),
“Towards Modeling Lean Blow Out in Gas Turbine Flameholder Applications,” GT2004-53967, Proceedings of ASME Turbo
Expo 2004, June 2004, Vienna, Austria.
Leong, M. Y., Hautman, D. J., (2002), “Near-Field Spray Characterization of a Liquid Fuel Jet Injected Into a Crossflow,”
ILASS Americas, 15th Annual Conference on Liquid Atomization and Spray Systems, Madison, WI, May 2002
Locke, R.J., Hicks, Y.R., Anderson, R.C., Zaller, M.M, (1998), “Optical Fuel Injector Patternation Measurements in
Advanced Liquid-Fueled, High Pressure, Gas Turbine Combustors,” Combustion Sci. and Tech., 1998, Vol. 138, pp. 297-311.
Law, C. K, “Mechanisms of Flame Stabilization in Subsonic and Supersonic Flows,” Major Research Topics in
Combustion, Eds.: M.Y.Hussaini, A.Kumar, R.G.Voigt, Springer-Verlag, pp. 201-236, 1992.
Madabhushi, R. K., Leong, M. Y., Hautman, D. J., “Simulation of the Break-up of a Liquid Jet in Crossflow at Atmospheric
Conditions,” GT2004-54093, Proceedings of ASME Turbo Expo 2004, June 2004, Vienna, Austria
Mehta, P.G., and Soteriou, M.C. (2003), “Combustion Heat Release Effects on the Dynamics of Bluff Body Stabilized
Premixed Reacting Flows,” AIAA Paper 2003-0835, January 2003
Mellor, A.M. (1980), “Semi-Emperical Correlations for Gas Turbine Emissions, Ignition, and Flame Stabilization,” Prog.
Energy Combustion Science, Vol. 6, pp. 347-358.
Menon, S. (2003), “Modeling Pollutant Emission and Lean Blow Out in Gas Turbine Combustors,” AIAA Paper 2003-4496,
Joint Propulsion Conference
Ozawa, R.I., “Survey of Basic Data on Flame Stabilization and Propagation for High Speed Combustion Systems,” AFAPL-
TR-70-81, November 1970.
Raffoul, C.N., Nejad, A.S., Gould, R.D., “Three-Component Velocity Measurements Downstream of a Bluff-Body
Flameholder,” Final Report, WL-TR-95-2120, 1995.

11
American Institute of Aeronautics and Astronautics
Shahnam, M., Wu, P.K., Kirkendall, K.A. (1998), “Combustion Instability of a Diffusion Flame Using an Integral Fuel
Injector/Flameholder Device,” AIAA Paper 98-0639, 36th Aerospace Sciences Meeting & Exhibit, Reno, NV, January 1998.
Spadaccini and TeVelde, “Autoignition Characteristics of Aircraft Type Fuels,” NASA CR-159886, 1971.
Sujith, R.I, Waldherr, G.A., Jagoda, J.I., Zinn, B.T., (1999) “A Theoretical Investigation of the Behavior of Droplets in Axial
Acoustic Fields,” J. of Vibration and Acoustics, vol. 121, July 1999, pp. 286-294.
Wadia, A. R., and James, F. D., “F110-GE-132: Enhanced Power Through Low-Risk Derivative Technology,” Journal of
Turbomachinery, Vol. 123, 2001, pp. 544-551.
Williams, F.A., “Flame Stabilization of Turbulent Premixed Gases,” pp. 1157-1170, 1955.
Wotel, G.J., Gallagher, K.E., “High-Speed Turboramjet Ramburner Technology Development,” 1989 JANNAF Propulsion
Meeting, Cleveland, OH, 1989.
Wu, Pei-Kuan, Kirkendall, K.A., Fuller, R.P, “Breakup Process of Liquid Jets in Subsonic Crossflows (1997), ” J. of
Propulsion and Power, Vol. 13, No. 1, pp. 64-73.
.

12
American Institute of Aeronautics and Astronautics

You might also like