Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Implications of Turbine Erosion

for an Aero-Engine’s
High-Pressure-Turbine Blade’s
Low-Cycle-Fatigue

Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Muhammad Naeem
Institute of Space Technology (IST),
Life-Consumption
P.O. Box 2750,
Islamabad 44000, Pakistan Some in-service deterioration in any mechanical device, such as a military aero-engine,
is inevitable. As a result of experiencing any deterioration, an aero-engine will seek a
different steady operating point thereby resulting in a variation in the high-pressure spool
speeds in order to provide the same thrust to keep aircraft’s performance invariant. Any
increase in the high-pressure spool speed results in greater low-cycle fatigue damage for
the hot-end components and thereby higher engine’s life-cycle costs. Possessing better
knowledge (of the impacts of high-pressure turbine’s erosion upon the low-cycle fatigue
life-consumption of aero-engine’s hot-end components) helps the users to take wiser
management decisions. For a military aircraft’s mission profile, using bespoke computer
simulations, the impacts of turbine erosion for high-pressure turbine-blade’s low-cycle
fatigue life-consumption have been predicted. 关DOI: 10.1115/1.3078383兴

1 Introduction and Background 2 Performance Deterioration of Gas Turbines


High-performance aircraft, as used in modern aviation, espe- In an ideal world, an engine would operate with the same per-
cially for military purposes, are complex in design and are re- formance from the time it enters service until it is removed. This
quired to operate under severe stresses and temperatures 关1兴. Thus of course does not happen in reality as the engine will deteriorate.
designer and users of these aircraft continually seek greater reli- There is very little reliable quantitative data on the magnitudes of
ability, increased availability, enhanced performance, and im- performance deterioration of engine components as their service
proved safety as well as low life-cycle costs. In-service costs con- lives are extended, except for that in the papers written by Grewal
sist mainly of those associated with 关2兴 共i兲 the fuel consumed 关7兴, Sallee and co-workers 关8,9兴, and Saravanmutto and co-
during the operation, and 共ii兲 the replacement of the system’s workers 关10,11兴. Sallee’s 关8兴 paper pertains to studies conducted
components. Therefore any extensions of life expectations or re- on the “Pratt and Whitney JT9D family of engines.” Sallee 关8兴
ductions in fuel-usage of aero-engines directly lower the life-cycle produced several performance deterioration mathematical models
costs and depend on the types of operation or mission undertaken, for the JT9D engine’s behavior. These models show that the
operating conditions experienced, and rate of in-service aero- performance-loss mechanisms are associated with the following.
engine deterioration. 共a兲 The compressors cause reductions in both flow capacity
Some in-service deterioration in any mechanical device, such as and efficiency.
a military aero-engine, is inevitable. Several publications describe 共b兲 The turbines result in a flow capacity increase and an
engine-performance deterioration. A preliminary investigation re- efficiency fall.
garding the impacts of aero-engine’s deteriorations on military
aircraft mission’s effectiveness established that there is a signifi-
cant adverse impact of aero-engine deteriorations. It also indicated
3 Erosion
that the extent of adverse impacts varies with the type of aero- This, in the present context, is the abrasive removal of material
engine’s component. Among the major aero-engine’s components, from the flow-path components by hard particles suspended in the
low-pressure compressor 共LPC兲 being the overall severest fol- gas stream. As a result, the gas-turbine’s aerofoil blades become
lowed by high-pressure turbine 共HPT兲 关3兴. This necessitated the eroded, some of the leading edges blunted, the trailing edges
need for a comprehensive investigation for the impacts of LPC’s thinned and the surface roughness increased. It also causes losses
fouling as well as of HPT’s erosion. Hence this investigation was of the blades’ camber and length, as well as of the seal material.
undertaken. The impacts of LPC’s fouling have already been in- Erosion of the aerofoil also occurs as a result of the engine’s
vestigated 关4–6兴. ingestion of foreign particles. The rate of foreign particle entry
There are many components in a gas-turbine engine, but its will be greatly influenced by the engine-intake design. This is
performance is highly sensitive to changes in only a few. The particularly evident with aircrafts such as the General Dynamics
majority of these potentially sensitive parts are the rotating com- F-16 whose intake acts as a large air scoop and is located below
ponents. Among these the HPT blades are the most sensitive com- the belly of the aircraft in contrast to the Russian MIG 29. When
the MIG 29 is on the ground, the main intakes are blocked off and
ponents, because they are subjected to both the highest rotating
flow is directed to the engines through open doors located on top
speeds and gas temperatures 关3兴 and so have been selected for
of the aircraft 关12兴; these doors are shut after take-off.
present investigation.
The erosion of aerofoil’s profile leads to changes in the aero-
foil’s inlet angle and throat opening. The consequent widening of
Manuscript received May 6, 2008; final manuscript received December 1, 2008; the tip and seal clearances results in increased air-leakage losses
published online May 22, 2009. Review conducted by David Walls. 关2兴.

Journal of Engineering for Gas Turbines and Power SEPTEMBER 2009, Vol. 131 / 052501-1
Copyright © 2009 by ASME
4 High-Pressure Turbine Performance-Deterioration aircrafts differ by an order of magnitude 关16兴. In addition, there is
evidence of large variations in throttle handling that may occur
The performance-loss mechanisms, which cause the majority of
between different pilots flying the same mission even using the
HPT deteriorations, are 共i兲 blade-tip clearance increases and 共ii兲
same aircraft type. May et al. 关17兴, after studying several United
vane twisting and bowing. Increased clearances between the rotor
States Air Force missions flown in F15, F5, and A10 aircrafts,
blades and the casing are caused predominately by centrifugal and
concluded that different pilots made significantly different usages
thermal loads imposed during engine transients and by distortions
of the throttle while flying the same mission. This led to a varia-
of the engine casing as a result of changing flight loads 关12兴. Vane
tion by a factor of as much as 5 in the number of idle-maximum
distortion ensues as a result of both aerodynamic loads and from
power-idle throttle-excursions per sortie. Consequently, efforts to
stresses due to thermal gradients. As the vanes distort, coolant air
develop LCF life-usage monitoring have primarily been focused

Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
is allowed into the main gas stream resulting in a reduction in the
on military operations in order to achieve economic lives in this
turbine’s efficiency 关2兴. Bowing of the HPT’s vanes causes the
highly-demanding environment without compromising safety.
flow area to increase and hence an increase in the flow capacity
关13兴.
8 Aircraft Gas-Turbine Engine LCF Life-
5 Potential-Failure Modes Consumption
The major ones for turbine components are creep, mechanical It is necessary to carry out two steps to ensure that the useful
fatigue, i.e., low-cycle fatigue 共LCF兲 and vibration induced high- lives of critical components are not exceeded. The first is the
cycle fatigue 共HCF兲, thermal fatigue, oxidation, and sulphidation. setting of an accurate life expectancy. The second is a means for
At temperatures below approximately 800° C, mechanical fatigue determining and controlling the rate of life consumption in actual
is usually the dominant failure mode. At higher temperatures, i.e., service.
above about 1000° C, creep, oxidation, and thermal fatigue 共acting The fatigue-damage calculation algorithm can be analyzed in
alone or together兲 usually cause failure. In the intermediate tem- the following three discrete steps 关2兴.
perature range, any of the several failure modes described can be 共1兲 Determine the number of load cycles already completed
dominant, depending on the structure, the material, and the engine from the measured values of the pertinent engine param-
cycle employed. Despite possible interactions of the major failure eters.
modes, a region of available life of the component does exit 关14兴. 共2兲 Calculate the fatigue damage incurred during each indi-
vidual load-cycle.
6 Low-Cycle Fatigue 共3兲 Estimate the total life consumed by summing up the dam-
This is caused by the imposition of varying loads 共and hence age incurred during each individual load-cycle.
stresses兲 on a component of an aero-engine. LCF occurs after
relatively low numbers 共typically ⬍50,000兲 of cycles of high- 9 Cycle Counting Methods
stress amplitude. Due to the high-stress levels implicit in LCF, it is LCF damage of an aero-engine results from the application of
normally accompanied by plastic deformation and is often termed varying loads. In order to determine these loads, it is necessary to
“high-strain fatigue,” as plastic strain is the primary parameter relate them to measurable engine parameters. For example, the
governing life. For a military-fighter aircraft’s engine, LCF is centrifugal stresses in a place can be determined 关18兴 via

冋 册
probably the most significant life-reducing failure-mechanism and
is the direct result of the many throttle changes experienced dur- ⌽i 2

ing this application. In general, the engine components that are Si = Sref 共1兲
⌽ref
subject to LCF are also those that are unlikely to be contained
共because of the normally very high rotational speeds and thereby An essential step in the prediction of fatigue life is the reduc-
damage occurring to components down stream兲 in the event of a tion of service strain or stress history to a series of cycles and
fatigue failure. It is these that are critical to flight safety and half-cycles. This process is known as the cycle-counting method.
should be monitored. Various methods have been developed to identify the individual
In the 1960s, most critical components were limited by creep cycles from a complex loading-history. One of the early cycle-
and stress rupture properties. Less than 1% of all rotating compo- extraction methods was the Range–Mean method. In an attempt to
nents were life limited by LCF. However, today’s designs find overcome the problems inherent in the Range–Mean and similar
well over 75% of these components limited in life by LCF 关15兴. counting methods, as discussed in more detail by Dowling 关19兴,
various alternative methods have been proposed, which share the
7 Factors Affecting LCF Life Usage in Military Avia- following two characteristics.
tion 共1兲 Each part of the overall history is counted only once.
There is a wider spectrum of usage for military aero-engines 共2兲 Smaller ranges are counted down to a predetermined
because military aircrafts fly a greater variety of mission profiles threshold.
than civil airliners. It is rare for a modern military plane to be
restricted to a single role. This, together with the need to maintain Probably the most widely-used method is the rain-flow, or the
crew proficiency in various aspects of their roles, means that Pagoda-roof, method first developed by Matsuishi and Endo 关20兴
many different mission profiles are flown regularly. and then subsequently by others 关19,21,22兴. For a more complete
The large throttle movements necessary to exploit fully the per- description of the rain flow method, the reader is referred to Refs.
formance of military aircraft also have a dramatic impact on LCF 关19,21兴. The same method was used for the present investigation.
life consumption. It should be remembered that the relationship
between applied stress and remaining life is logarithmic, and that 10 Low-Cycle Fatigue-Damage Calculation
mechanical stress imposed on the engine varies as the square of its The chosen cycle-counting technique will transform the com-
rotational speed. Therefore, a modest increase in the range of plex load history into a succession of stress ranges about a mean
throttle movement can have a drastic impact on the fatigue-life stress. The next stage in the algorithm process is to calculate the
consumed. actual fatigue damage caused by each individual cycle.
There can also be huge variations in the engine usage on what Two important methods for calculation of LCF damage are the
are nominally-identical missions. In formation flying, the throttle stress-life and strain-life methods 关2兴. However, due to the con-
usages for the lead aircraft and the one at the rear of the formation cept of LCF as high-strain fatigue, the strain-life method is gen-
and also the LCF life usage between the engines of these two erally considered to be more appropriate. First, the nominal cen-

052501-2 / Vol. 131, SEPTEMBER 2009 Transactions of the ASME


trifugal stress is calculated at each measured value of spool speed environment, etc., on engine performance can be studied and ap-
using Eq. 共1兲. Then the local stress and strain at the notch for the preciated before they are encountered in service. The more accu-
first applied nominal stress is calculated using Neuber’s rule 共Eq. rate the simulation, i.e., preferably being based on actual mea-
共2兲兲 and the cyclic stress-strain curve 共Eq. 共3兲兲, which may be sured data, the more realistic will be the predictions achieved.
combined to give Eq. 共4兲 关23兴, as follows: For the purpose of the present investigation the “Engine’s
Performance-Simulation Program 共NAEEMPAK兲,” “Aircraft’s
␴␧
Kt2 = 共2兲 Flight-Path and Performance-Simulation Program,” “Aircraft and
S e⬘ Engine’s Performance-Simulation Program 共NAEEMPAKA兲,” and

冋 册
“Low-Cycle Fatigue Life-Usage Prediction Program 共NAEEM-
1/n⬘
␴ ␴ PAKB兲” have been used. These programs were developed by the

Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
␧= + 共3兲 author while completing his Ph.D. at the Cranfield Unversity, UK,
E K⬘
and are proprietary to the author. The programs were further en-

冋册 冋 册
␴2
E
+␴

K⬘
1/n⬘
=
共KtS兲2
E
共4兲
hanced and tailored to accomplish the present task.
NAEEMPAK is basically an updated version of the engine-
performance simulation program TURBOMATCH. TURBOMATCH has
For subsequent load-applications, the local delta stress is calcu- been used widely at the Cranfield University for many years. It
lated by combining Eq. 共2兲 with the hysteris stress-strain rule of facilitates design point and off-design performance calculations
Eq. 共5兲 to give Eq. 共6兲 关23兴. and has the ability to simulate the behavior of an engine and the

冋 册
effects of implanted faults in its components, for instance, a re-
1/n⬘
⌬␧ ⌬␴ ⌬␴ duction in component efficiency or mass flow capacity. It is based
= + 共5兲
2 2E 2K⬘ on the aero- and thermodynamic balancing of the gas turbines,
using the method of component matching in which the compat-
共⌬␴兲2
2E
+ ⌬␴
⌬␴
冋 册
2K⬘
1/n⬘
=
共Kt⌬S兲2
2E
共6兲
ibilities of the flow between the various compressors and turbines
are determined.
The Aircraft’s Flight-Path and Performance-Simulation Pro-
The local notch stress can be deduced by adding ⌬␴ to the gram integrates the simulation of each flight stage to give a com-
value of ␴ at the origin of the hysteris loop, and the local strain plete mission profile for the military aircraft. Details regarding
range ⌬␧ can be determined from aircraft’s performance throughout the mission profile, as described
in referred publications 关2,27,28兴, formed the basis for develop-
共Kt⌬S兲2 ment of this program.
⌬␧ = 共7兲
⌬␴E NAEEMPAKA is basically a combination of NAEEMPAK and the

This value of local strain range is then inserted into the strain- aircraft’s flight path and performance-simulation program. This
life Eq. 共8兲 to determine the cyclic life N f , as follows: program simulates the flight path of the military aircraft. The NAE-
EMPAK supplies the engine’s net thrust and specific fuel-
⌬␧ ⌬␧e ⌬␧ p ␴ f consumption values to the aircraft’s flight path and performance-
= + = 共2N f 兲b + ␧ f 共2N f 兲c 共8兲 simulation program, so enabling the pertinent flight parameters to
2 2 2 E
be evaluated for each flight-stage. It requires data for 共i兲 the en-
Equation 共8兲 can be modified to account for the mean stress gines, 共ii兲 the geometric and aerodynamic characteristics of the
effects by including the mean stress ␴0 in the elastic strain term as aircraft, and 共iii兲 the mission profile. These inputs are defined by
proposed by Morrow 关24兴. This leads to the user through separate input files.
⌬␧ 共␴ f − ␴0兲 NAEEMPAKB has been used for the purpose of LCF life evalua-
= 共2N f 兲b + ␧ f 共2N f 兲c 共9兲 tion 关2兴. The strain-life technique 共as described 共very briefly兲 ear-
2 E lier兲 has formed the basis for the development of this computer
The final step is to sum up the damage arising during each program. Program requires data for 共i兲 the engine’s design and
individual cycle. Several researchers have presented theoretical blade material 共defined by the user through a separate file兲 and 共ii兲
syntheses for predicting the cumulative damage. The most com- high-pressure spool speed 共HPSS兲 history 共obtained from the NAE-
mon and well known is the Palmgren–Minor cumulative damage EMPAKA兲.
law 关25兴 expressed as
12 HPT Blade’s Material
n f n1 n2 n3
= + + +¯ 共10兲 Despite a wide literature review the author was unable to de-
N f N1 N2 N3
termine the exact type and specification of the material used for
When the value n f / N f , becomes unity, the component is con- HPT’s blades of the F404 engine. Therefore, it was decided to
sidered to have failed. In reality, component failure occurs at ra- assume any of the typical materials 共such as MAR M002, INCO
tios of between 0.61 and 1.49 关26兴. As pointed out in the ESDU 792, INCO 713, INCO 718, NIMONIC 75, NIMONIC 80A, NI-
paper on fatigue-life estimation 关23兴, although the Palmgren– MONIC 115, etc.兲 used for manufacturing of turbine’s blade. Dur-
Minor law may be considered inaccurate in some respects, it is ing literature review to determine the material properties required
simple and has been widely applied, and the reasons for its inad- for predicting the LCF life 共for the method chosen兲, the author
equacies are recognized and can be compensated for in the design could only find the relevant data for INCO 718, hence this mate-
process. rial was used for the present investigation. The relevant LCF prop-
erties for INCO 718 are shown in Table 1.
11 Computer Modeling and Simulations
Because of the enormous cost reductions and rapid results 13 Aircraft and Engines Chosen
achieved relative to experimental/trial techniques, the use of vali- For this investigation, the performance of a McDonnell Dou-
dated computer-simulation techniques has attained recently the glas F-18 aircraft, powered by two nominally-identical F404-GE-
status of an advanced engineering procedure. Many scenarios can 400 aero-engines, has been assessed. The reason for selecting the
be simulated without incurring the major difficulties and expenses F404 was twofold. First, the F404 engine is a typical new-
of preparing and testing engines. Hence the implications of vari- technology engine, and, as such, the results of the investigation
ous factors such as variations in deteriorations for engine as a will be applicable to other similar new-generation engines. Sec-
whole or either of its major components, changes in operating ond, sufficient relevant data were available.

Journal of Engineering for Gas Turbines and Power SEPTEMBER 2009, Vol. 131 / 052501-3
Table 1 Fatigue properties for the INCO 718 search 关2兴. Hence, for this assessment, a single-term empirical
“erosion index 共EI兲” was defined to describe the effect of the
Fatigue property Numerical value reduction in efficiency as well as the increase in mass flow capac-
ity of HPT. For example, for an erosion model 共EM兲 1:0.5 共i.e.,
Young’s modulus 158⫻ 109
HPT’s efficiency to mass flow capacity deteriorations’ ratio兲, an
Cyclic strength coefficient 1433⫻ 106
Cyclic strain coefficient 0.0945 efficiency of 96% and a flow capacity of 102% of those for a
Fatigue strength coefficient 1311⫻ 106 clean engine may be described as 4% EI. As the erosion usually
Fatigue strength exponent ⫺0.0596 starts from irregular pits and crevices randomly distributed on the
Fatigue ductility coefficient 0.389 blade contour, its effects on the turbine performance are insignifi-
Fatigue ductility exponent ⫺0.631 cant initially. Therefore more severe erosion damages, which no-

Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Endurance strain 0.001 ticeably influence enough the turbine performance, have been
considered. As such, for the purpose of the present analysis, the
effects of EM 1:0.5 for the following set of engine conditions
have been assessed:
14 Assumed Mission-Profile
• clean engines 共i.e., a HPT’s EI of zero for the both engines兲
In general, unlike those for civil aircraft, the mission profiles • HPT’s EI of 1%, 2%, 3%, … or 6% for the both engines
for military aircrafts can be relatively complex. However, for the
purpose of the present analysis, the following 共a relatively simple兲 The effects of varying proportions of HPT’s mass flow capacity
aircraft mission profile has been assumed: deterioration for a given HPT’s efficiency deterioration level have
been also assessed by considering EMs 1:0.25, 1:0.5, 1:0.75, 1:1,
共1兲 short take-off 共TO兲 with reheat 共RH兲 on, followed by
1:1.5, 1:2, 1:3, and 1:4.
climbing after TO, from sea level to 6000 m 共m兲, while
For the purpose of the present analysis, HPT’s EI of 6% is
accelerating to Mach number 共M兲 0.6 within 420 s 共s兲
considered an outage for military operations, i.e., the breaking
共2兲 climbing from 6000 m to 8000 m while accelerating from
down of a component and thereby stopping the further use of
M 0.6 to 0.7 in 250 s
machine 共aero-engine in this case兲 until an appropriate replace-
共3兲 climbing at a constant M 0.7, from 8000 m to 10,500 m in
ment or repair is undertaken 关2兴.
100 s, followed by acceleration 共ACC兲 from M 0.7 to 1.0
within 100 s
共4兲 cruising at 10,500 m and M 1.0 until a preset time 共1800 s 16 Discussions and Analysis of Results
from take-off兲 followed by deceleration to M 0.95 within The aim of the present investigation is to quantify the impacts
60 s of a military aircraft’s turbofan HPT’s erosion on the blade’s LCF
共5兲 enhancement of engines’ power to 80%, attaining and then life-consumption for a specified mission profile. An important fac-
cruising at maximum M 共for total time duration of 900 s兲 tor, which needs to be realized when examining the results, is that
followed by acceleration/deceleration to M 0.85 within 60 s the numerical values for the effects of HPT’s erosion are not de-
共6兲 cruising at 10,500 m and M 0.85 for a set time period of finitive but nevertheless are considered good “ball park” figures.
900 s followed by cruising to cover a distance of 500 km In addition, the trends of the effects are important as they provide
共km兲 insight into how the engine behaves.
共7兲 accelerating as a result of RH for 60 s, followed by decel- The prime factor responsible for any change in the HPT blade’s
erating by switching off the RH, to M 0.85 within 200 s LCF life-consumption is the variation in the HPSS. A higher
共8兲 descending at a constant M 0.85, from 10,500 m to 8000 m HPSS results in greater LCF damage and thereby higher blade’s
within 300 s, followed by cruising toward a set target at LCF life-consumption. The variations in HPT’s erosion levels
3000 km from home base would require the engine共s兲 to run at different HPSS and/or tur-
共9兲 descending from 8000 m to 4500 m while decelerating to bine entry-temperatures 共TETs兲 in order to meet the thrust require-
M 0.5 within 250 s, followed by loitering for 100 s ment for achieving the same aircraft’s performance.
共10兲 descending to land within 350 s, followed by landing ap-
proach, flare, touch down, ground roll, and finally 17 Aero-Engine’s Behavior
switch-off
Any mission profile would always be a combination of altitude,
M, and engine power setting. Therefore, prior to an overall analy-
15 Undertaking the Computer Simulation sis of HPT blade’s LCF life-consumption, it was considered ap-
The basic methodology is to fly the aircraft through a complete propriate to first see how the concerned engine parameter is af-
mission-profile with 共i兲 both engines functioning properly and 共ii兲 fected by any variation in altitude/M/day temperature/engine
with both engines suffering identical prescribed degrees of HPT’s power setting or any combination of these. For the purpose of the
erosion. Subsequently the NAEEMPAKB is run using the results of present investigation, day temperature variations have been ex-
the NAEEMPAKA along with relevant material data as inputs. Thus pressed in terms of the international standard atmosphere devia-
the LCF life usage is predicted for each flight, so determining the tion 共ISAD兲 共i.e., an ISAD of 0 K for a standard day temperature
impacts of the HPT’s erosion on the LCF life usage as a function of 15° C and ISAD of ⫺15 K and 35 K for day temperatures of
of the aircraft’s flight path. 0 ° C and 50° C, respectively兲. All HPSS and maximum high-
The relationship between physical degradation and simulated pressure spool speed 共MHPSS兲 values have been expressed as
degradation is realized by choosing certain ratios between compo- percentage of its design point value. The results are analyzed in
nent efficiency and mass flow capacity degradations. Sallee and subsequent paragraphs.
co-workers 关8,9兴 described how the efficiency and mass flow ca- Figures 1–7 illustrate the variation in HPSS with increasing
pacity are affected by the engine configuration and degradation. altitude/M/ISAD/TET/HPT’s erosion, and at constant LPSS/NT
Although several publications have described different HPT’s ef- conditions 共with increasing HPT’s erosion兲 for stipulated HPT’s
ficiencies to mass flow capacity deteriorations ratios, as yet, pre- EMs, and with increasing HPT’s flow capacities for a given HPT’s
cise performance-parameters change due to typical faults in the efficiency level, respectively. A critical analysis establishes the
engine’s components and their inter-relationships are not known significant and clear variation trends of HPSS for different sets of
accurately 关29兴. The exact ratios chosen can be open to some stipulated conditions.
individual subjectivity and so it was worthwhile establishing the With increasing altitude, for the same levels of M, TET, and
values of their ratios commonly concluded from the previous re- ISAD, the HPSS increases very slightly but linearly within the

052501-4 / Vol. 131, SEPTEMBER 2009 Transactions of the ASME


Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Fig. 1 HPSS at M 0.95, TET 1555 K, and ISAD 0 K for stipulated Fig. 4 HPSS at an altitude 10,000 m, M 0.95, and ISAD 0 K for
HPT EMs with increasing altitude stipulated HPT EMs with increasing TET

With increasing ISAD, for the same levels of altitude, M, and


stratosphere. However, within the troposphere it varies signifi- TET, the HPSS initially increases almost linearly followed by a
cantly following a clear variation trend 共i.e., an almost linear in- decreasing rate until reaching its peak value and subsequently
crease until reaching its peak at about 6000 m altitude and subse- reduces with slightly increasing rate 共see Fig. 3兲. With an increase
quently an almost linear reduction until the end of the in altitude/TET, the trends of variation remain almost the same but
troposphere兲 共see Fig. 1兲. With an increase in M/ISAD, the trend whole pattern drifts toward the right 共i.e., in the direction of in-
in variations within the troposphere remains almost the same but creasing ISAD兲, whereas with an increase in M, it drifts toward
the whole pattern drifts toward the right, i.e., in the direction of left 关30兴.
increasing altitude, whereas with an increase in TET, it drifts to- With increasing TET, for the same levels of altitude, M, and
ward left. At a given altitude, HPSS is significantly less at lower ISAD, the HPSS increases significantly and with slightly decreas-
TET levels 关30兴. ing rate 共see Fig. 4兲. At lower TET levels, the HPSS at lower
With increasing M, for the same levels of altitude, TET and ISAD/M, at higher altitude, and for engines without RH are
ISAD, the HPSS initially increases with slightly increasing rate, higher, whereas at higher TET levels, the HPSS at higher
then linearly followed by a decreasing rate reaching its peak at ISAD/M, at lower altitude, and for engines with RH are higher.
about M 1.3 and subsequently reduces almost linearly 共see Fig. 2兲. HPSS is higher for engines without RH at TET levels lower than
With an increase in altitude/TET, the whole pattern drifts toward the design point, whereas it is lower for engines without RH at
the right 共i.e., in the direction of increasing M兲, whereas with an TET levels higher than the design point 关30兴.
increase in ISAD, it drifts toward left 关30兴.

Fig. 5 HPSS at an altitude 10,000 m, M 0.95, TET 1555 K, and


Fig. 2 HPSS at an altitude 10,000 m, TET 1555 K, and ISAD 0 K ISAD 0 K for stipulated HPT EMs with increasing HPT’s erosion
for stipulated HPT EMs with increasing M

Fig. 6 HPSS at an altitude 10,000 m, M 0.95, ISAD 0 K, and


Fig. 3 HPSS at an altitude 10,000 m, M 0.95, and TET 1555 K constant NT for stipulated HPT EMs with increasing HPT’s
for stipulated HPT EMs with increasing ISAD erosion

Journal of Engineering for Gas Turbines and Power SEPTEMBER 2009, Vol. 131 / 052501-5
Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Fig. 7 HPSS at an altitude 10,000 m, M 0.95, ISAD 0 K, and a Fig. 9 HPSS at stipulated conditions with increasing HPT’s
constant LPSS for stipulated HPT EMs with increasing HPT’s erosion
erosion

phases for a given EM and HPT’s erosion; 共ii兲 HPSS at the start
With increasing HPT’s erosion, for the same levels of altitude, and end of TO, RH, and ACC phases for a given EM but with
M, TET, and ISAD, the HPSS decreases significantly and with increasing HPT’s erosion; and 共iii兲 MHPSS during TO, ACC, RH,
slightly increasing rate 共see Fig. 5兲. and enhancement of power to 80% followed by cruising at con-
With increasing HPT’s erosion, for the same levels of altitude, stant power setting flight phases for the stipulated EM but with
M, and ISAD, the HPSS decreases with slightly increasing rate for increasing HPT’s erosion, respectively. A critical analysis estab-
constant net thrust 共NT兲 and low-pressure spool speed 共LPSS兲 lishes a significant variation 共with differing variation trends兲 in
conditions considered separately 共see Figs. 6 and 7兲. However, HPSS/MHPSS during different flight phases as well as for differ-
this is accompanied with significant variations in TET. ent points on the mission profile with varying EMs and HPT’s
At the same conditions, HPSS is significantly higher for higher erosion levels.
HPT’s flow capacities for a given HPT’s efficiency level 共see Figs. For the conditions assumed for the present investigation, for a
1–7兲. However, with increasing flow capacity levels, its peak given HPT’s EM and with the same HPT’s erosion levels, HPSS
value slightly drifts toward the left with increasing altitude, 共i兲 increases initially with a decreasing rate, however, subse-
whereas it slightly drifts toward the right with increasing M/ISAD quently remains almost invariant during TO; 共ii兲 increases with
共see Figs. 1–3兲. slightly increasing rate during RH; and 共iii兲 reduces initially with
a slightly decreasing rate, reaches minimum value, and then starts
18 Aero-Engine: Aircraft Combination’s Behavior increasing with slightly increasing rate 共i.e., there is a dip兲 during
ACC 共see Fig. 8兲.
For the purpose of subject investigation, while the aircraft flies
through the assumed mission profile, in addition to the behavior of
fitted aero-engines in isolation, the behavior of aero-engines’ and
aircraft’s aerodynamic structure combination as a whole is of
greater importance. During flight segments such as TO and
combat/RH phases, the aero-engines’ peak performance plays the
dominant role and the aircraft performs according to the maxi-
mum available NT from the engines. Whereas during some other
flight segments, e.g., cruising at a specified altitude and M, the
aircraft’s aerodynamic characteristics play the dominant role and
aero-engines are run at a power setting such that the NT available
from the engines balances the aircraft’s drag. Therefore, the en-
gine parameter of concern, i.e., HPSS/MHPSS, has been also ob-
served during some important flight segments and results analyzed
in subsequent paragraphs. Finally the aero-engine’s HPT blade’s
LCF life-consumption has been predicted for the complete mis-
sion profile. Fig. 10 MHPSS during TO for stipulated EMs with increasing
Figures 8–13 illustrate 共i兲 HPSS during TO, RH, and ACC HPT’s erosion

Fig. 11 MHPSS during ACC for stipulated EMs with increasing


Fig. 8 HPSS during stipulated flight phase HPT’s erosion

052501-6 / Vol. 131, SEPTEMBER 2009 Transactions of the ASME


Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Fig. 12 MHPSS during RH for stipulated EMs with increasing Fig. 14 HPT blade’s LCF life-consumption „for stipulated mis-
HPT’s erosion sion profiles… with increasing HPT’s erosion

At any given point, with increasing HPT’s erosion, HPSS 共i兲 distinct variation trend of MHPSS is also observed for “enhance-
reduces almost linearly during RH; 共ii兲 reduces almost linearly ment of engines’ power to 80%, followed by cruising at constant
共and with a slightly decreasing rate from start to end of TO兲 dur- power setting” flight phase 共see Fig. 13兲.
ing TO; and 共iii兲 reduces almost linearly and with slightly increas-
ing rate in earlier and later parts of acceleration, respectively 共see
Fig. 9兲. For example at the start of TO, it is 100%, 98.8%, 97.6%, 19 Representative Peacetime Training Mission Profiles
and 96.4% with HPT’s EM 共1, 0.50兲 and EI of 0%, 2%, 4%. and for a Military Aircraft
6%, respectively, i.e., a reduction of 3.6% as a result of an in- The details covered so far establish the significance of varia-
crease in HPT’s EI from 0% to 6%. Whereas at the end of TO, it tions in altitude, M, engine power settings, etc., for HPSS of aero-
is 100%, 99.1%, 98%, and 96.8%, i.e., a reduction of 3.2%, re- engine of a military aircraft. Also typical mission profiles for a
spectively. military aircraft vary significantly 共dependent on their role兲 in
With increasing HPT’s erosion, for various EMs, a clear varia- terms of altitude, M, and required engine power settings. There-
tion trend is observed for MHPSS during TO. At lower flow ca- fore, instead of restricting the present investigation to the assumed
pacity levels for a given efficiency level, e.g., EM 共1, 0.25兲, the mission profile, it was considered appropriate to extend the analy-
MHPSS reduces almost linearly with increasing HPT’s EI. ses by considering representative peacetime training mission pro-
Whereas with increased flow capacities, the reduction in MHPSS files such as air superiority, simple cruise, low-level strike, and
no more remains linear, rather it reduces with slightly increasing strategic bombing for a military aircraft 关5兴.
rate. However, at any given EI, it is significantly higher for higher
flow capacities for a given HPT’s efficiency level. Also with in-
creasing flow capacities, the extent of reduction in MHPSS re- 20 HPT Blade’s LCF Life Consumption
duces with increasing EI. As a result, for an EM 共1, 1.5兲, the For a given mission profile, an overall increase in HPSS as a
MHPSS remains almost invariant from an EI of 0% to 2% result of HPT’s erosion leads to higher LCF damage and thereby
whereas for an EM 共1, 4兲, MHPSS even increases from 100% to affects the blade’s LCF life-consumption adversely, i.e., LCF life-
100.85% for an increase in EI from 0% to 3% and subsequently consumption increases. For the conditions assumed for the present
MHPSS starts decreasing with slightly increasing rate 共see Fig. investigation 共i.e., for a LPSS limiting of 100% of its design point
10兲. value and no TET limiting兲 and for the same EMs, blades’ LCF
An almost similar variation trend is observed for MHPSS dur- life-consumption reduces significantly and with slightly decreas-
ing TO, ACC, and RH phases with only one exception. Unlike TO ing rate with increasing HPT’s erosion levels. For example for the
and RH phases where at lower flow capacity level for a given assumed mission profile 共AMP兲, for HPT’s EM 共1, 0.50兲, the
efficiency level, e.g., EM 共1, 0.25兲, the MHPSS reduces almost blade’s LCF life consumption with HPT’s EI of 2%, 4%, and 6%
linearly; during ACC phase, it reduces with slightly increasing is 7.59⫻ 10−4, 6.38⫻ 10−4, and 5.38⫻ 10−4, respectively, as com-
rate 共see Figs. 10–12兲. Like TO, ACC, and RH phases, a clear and pared with 7.92⫻ 10−4 for clean engines 共i.e., a reduction of
4.17%, 19.44%, and 32.07%, respectively兲. With increasing
HPT’s erosion level, the rate of reduction in blades’ LCF life-
consumption is almost the same for all mission profiles except
AMP where it is significantly higher 共see Fig. 14兲.
Among the four representative mission profiles of a military
aircraft, air superiority is the severest in terms of blade’s LCF life
consumption whereas the simple cruise is the least severe. At
HPT’s EI of 6%, blade’s LCF life consumption for air superiority
mission profile is 28.78%, 67.02%, and 71.61% higher as com-
pared with strategic bombing, low-level strike, and simple cruise
mission profiles, respectively. The significant variation in the lev-
els of blades’ LCF life-consumption for different mission profiles
is due to the reason that it varies with any change in nature/extent
of flight segment共s兲 as well as their interarrangement in any given
mission profile.
As a result of the reduced blade’s LCF life consumption, before
the HPT blades’ are considered unsafe for further use, for HPT’s
Fig. 13 MHPSS during enhancement of power to 80%, attain- EI of 6%, the aircraft can fly through air superiority, strategic
ing maximum M and cruising at constant power setting for bombing, low-level strike, and simple cruise mission profiles
stipulated EMs with increasing HPT’s erosion 1600, 2061, 2568, and 2603 times as compared with 1242, 1505,

Journal of Engineering for Gas Turbines and Power SEPTEMBER 2009, Vol. 131 / 052501-7
1909, and 1928 for HPT’s EI of 0%, respectively 共i.e., an increase References
of 28.82%, 36.94%, 34.52%, and 35.01%, respectively, as a result 关1兴 Devereux, B., and Singh, R., 1994, “Use of Computer Simulation Techniques
of an increase in HPT’s EI from 0% to 6%兲. to Assess Thrust Rating as a Means of Reducing Turbo-Jet Life Cycle Costs,”
For different conditions and flight segments discussed earlier, International Gas-Turbine and Aero Engine Congress and Exposition, The
the variation in HPSS/MHPSS for various EMs with increasing Hague, The Netherlands, Jun. 13–16.
关2兴 Naeem, M., 1999, “Implications of Aero-Engine Deterioration for a Military
HPT’s erosion levels is accompanied with significant variation in Aircraft’s Performance,” Ph.D. thesis, Cranfield University, UK.
TETs. For example at constant NT condition, and with EM 共1, 关3兴 Naeem, M., Singh, R., and Probert, D., 2001, “Impacts of Aero-Engine Dete-
0.50兲, TET for HPT’s EI of 6% increases to 1628 K as compared rioration on Military Aircraft Mission’s Effectiveness,” Aeronaut. J.,
with that of 1555 K for clean engines. An increase in TET leads to 105共1054兲, pp. 685–695.
关4兴 Naeem, M., 2006, “Impacts of Low-Pressure 共LP兲 Compressor’s Deterioration
greater creep and thermal fatigue damages of HPT’s blades 共areas

Downloaded from https://asmedigitalcollection.asme.org/gasturbinespower/article-pdf/131/5/052501/5071814/052501_1.pdf by Gas Turbine Research Establishment (GTRE) user on 24 September 2019
Upon an Aero-Engine’s High-Pressure 共HP兲 Turbine Blade’s Life Consump-
of future investigations兲. tion,” Aeronaut. J., 110共1106兲, pp. 227–238.
关5兴 Naeem, M., 2008, “Impacts of Low-Pressure 共LP兲 Compressor’s Deterioration
21 Conclusions of a Turbofan Engine Upon Fuel Usage of a Military Aircraft,” Aeronaut. J.,
112共1127兲, pp. 33–45.
Bespoke computer simulations have been used to explore the 关6兴 Naeem, M., 2008, “Impacts of Low-Pressure 共LP兲 Compressor’s Fouling of a
relative implications of turbine erosion for an aero-engine’s HPT- Turbofan Upon Operational-Effectiveness of a Military Aircraft,” Appl. En-
blade’s LCF life-consumption. The turbine erosion has a signifi- ergy, 85共4兲, pp. 243–270.
cant and favorable effect on the blade’s LCF life-consumption. 关7兴 Grewal, M. S., 1988, “Gas-Turbine Engine Performance Deterioration Model-
ling and Analysis,” Ph.D. thesis, Cranfield University, UK.
However, this is on the cost of a significant increase in TETs, 关8兴 Sallee, G. P., 1980, “Performance Deterioration Based on Existing 共Historical兲
thereby an expected significant increase in levels of creep and Data,” Report No. NASA-CR-135448.
thermal fatigue damages of HPT’s blades. The blade’s LCF life- 关9兴 Sallee, G. P., Kruckenberg, H. D., Toomy, E. H., 1986, “Analysis of Turbofan-
consumption varies significantly with the type of mission flown Engine Performance Deterioration and Proposed Follow-On Test,” Report No.
NASA-CR-134769.
and is the severest for air superiority mission profile. 关10兴 Saravanamuttoo, H. I. H., and Lakshminarasimha, A. N., 1985, “A Preliminary
The results of such an investigation as the present for a variety Assessment of Compressor Fouling,” Paper No. ASME 85-GT-153.
of mission profiles under various operating conditions/engine’s 关11兴 Zhu, P., and Saravanamuttoo, H. I. H., 1991, “Simulation of Advanced Twin-
deterioration levels will provide a wide database and can lead to Spool Industrial Gas Turbine,” ASME Paper No. 91-GT-34.
managers taking wiser decisions. Hence the engines’ utilization 关12兴 Little, P. D., 1994, “The Effects of Gas Turbine Engine Degradation on Life
Usage,” MS thesis, Cranfield University, UK.
and operational-effectiveness of the mission can usually be im- 关13兴 MacDonald, S., 1993, “A Dynamic Simulation of the GE F404 Engine for the
proved. However, greater benefits will be achieved if the present Purpose of Engine Health Monitoring,” MS thesis, Cranfield University, UK.
analysis is considered as an integral part of a more comprehensive 关14兴 Haub, G. L., and Hauhe, W. E., 1990, “Field Evaluation of On-Line Compres-
analysis including the aero-engine’s creep and thermal fatigue life sor Cleaning in Heavy-Duty Industrial Gas-Turbines,” ASME Paper No. 90-
GT-107.
usages, fuel-usage/weapon carrying capability, aircraft’s mission 关15兴 Nicholas, T., Haritos, G. K., and Christoff, J. R., 1985, “Evaluation of Cumu-
operational-effectiveness, and life cycle costs aspects. The de- lative Damage Models for Fatigue Crack Growth in an Aircraft Engine Alloy,”
tailed investigations regarding the impacts of turbine erosion on Jet Propuls., 1共2兲, pp. 131–136.
aero-engine’s HPT blades’ creep as well as thermal fatigue lives 关16兴 O’Connor, C. M., 1988, “Military Engine Condition Monitoring Systems: The
are underway. The results of analyses and predictions would be UK experience,” Proceedings of the Advisory Group for Aerospace Research
and Development, Neuilly Sur Seine, France, Paper No. AGARD-CP-448.
presented in the form of separate papers in near future. 关17兴 May, R. J., Jr., Chaffee, D. R., Stumbo, P. B., and Reitz, M. D., 1981, “Tactical
Aircraft Engine Usage: A Statistical Study,” American Institute of Aeronautics
Nomenclature and Astronautics, Report No. AIAA-81-1652.
关18兴 Balderstone, A. W., 1996, “A Generic Computer Model to Predict an Aero-
b , c ⫽ fatigue-strength and ductility exponent, Engine’s Low-Cycle Fatigue,” MS thesis, Cranfield University, UK.
respectively 关19兴 Dowling, N. E., 1972, “Fatigue Failure Predictions for Complicated Stress-
e⬘ ⫽ nominal strain Strain Histories,” J. Mater., 7共1兲, pp. 71–87.
关20兴 Matsuishi, M., and Endo, T., 1968, “Fatigue of Metals Subjected to Varying
E ⫽ elastic modulus
Stress,” Japan Society of Mechanical Engineers Conference, Fukuoka, Japan.
K⬘ ⫽ cyclic strength coefficient 关21兴 Downing, S. D., and Socie, D. F., 1982, “Simple Rainflow Counting Algo-
Kt ⫽ theoretical elastic-stress concentration factor rithms,” Int. J. Fatigue, 4共1兲, pp. 31–40.
M ⫽ free stream Mach number 关22兴 Rychlik, I., 1987, “A New Definition of the Rainflow Cycle Counting
N f ⫽ number of cycles to failure 共i.e., to the onset Method,” Int. J. Fatigue, 9共2兲, pp. 119–121.
关23兴 ESDU, 1995, “Fatigue-life Estimation Under Variable-Amplitude Loading Us-
of crack-initiation life兲 ing Cumulative Damage Calculations,” Fatigue-Endurance Data Contents,
Ni ⫽ number of stress cycles to failure under condi- Vol. 2, ESDU International, London.
tion i 关24兴 Morrow, J., 1968, Fatigue-Design Hand Book: A Guide for Product Design
ni ⫽ number of cycles applied at condition i and Development Engineers, Vol. 4, Society of Automotive Engineers, War-
rendale, PA.
n⬘ ⫽ cyclic strain hardening coefficient 关25兴 Palmgren, A., 1924, “Die lebensdauer von Kugellagern,” VDI Z. 共1857-1968兲,
S ⫽ stress imposed 68, pp. 339–341.
␧, ␧ f ⫽ local notch root strain and fatigue-ductility 关26兴 Miner, M. A., 1945, “Cumulative Damage in Fatigue,” ASME J. Appl. Mech.,
coefficient, respectively 12, pp. A159–A164.
关27兴 Eshelby, M., 1996, “Propulsion System Performance and Integration” Lecture
⌽i, ⌽ref ⫽ spool-speed at ith and reference condition, Notes, Department of Propulsion, Power, Energy and Automotive Engineering,
respectively School of Mechanical Engineering, Cranfield University, UK.
␴, ␴ f ⫽ local notch root stress and fatigue-strength co- 关28兴 Raymer, D. P., 1992, Aircraft Design: A Conceptual Approach 共AIAA Educa-
efficient, respectively tion Series兲,” American Institute of Aeronautics and Astronautics, Inc., Wash-
ington, DC..
␴0 ⫽ mean stress 关29兴 Stevenson, J. D., and Saravanamuttoo, H. I. H., 1995, “Effects of Cycle Choice
⌬S ⫽ blade stress range for cycle from blade speed and Deterioration on Thrust Indicators for Civil Engine,” 12th International
⌽L → ⌽H → ⌽L Symposium on Air-Breathing Engines, Melbourne, Australia, Sept. 10–15, Pa-
per No. ISABE 95-7077.
⌬␧, ⌬␧e, and ⌬␧ p ⫽ strain range, elastic strain range, and 关30兴 Naeem, M., 2008, “Implications of Day Temperature for A High-Pressure-
plastic strain range, respectively Turbine Blade’s Low-Cycle-Fatigue Life Consumption,” J. Propul. Power,
⌬␴ ⫽ local stress range 24共3兲, pp. 624–628.

052501-8 / Vol. 131, SEPTEMBER 2009 Transactions of the ASME

You might also like