Sei sulla pagina 1di 13

Engineering Failure Analysis 35 (2013) 33–45

Contents lists available at SciVerse ScienceDirect

Engineering Failure Analysis


journal homepage: www.elsevier.com/locate/engfailanal

Fracture mechanics in aircraft failure analysis: Uses and


limitations
R.J.H. Wanhill a,⇑, L. Molent b, S.A. Barter b
a
National Aerospace Laboratory NLR, Amsterdam, The Netherlands
b
Defence Science and Technology Organisation (DSTO), Melbourne, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Fracture mechanics is generally accepted as essential to the design of metallic aircraft
Available online 16 November 2012 structures, notably for predicting and validating the fatigue crack growth and residual
strength properties. Applications of Linear Elastic Fracture Mechanics (LEFM) began in
Keywords: the early 1970s, and would be expected by now to have reached a mature status. Hence
Fracture mechanics one might also expect that LEFM methods can often be used for analysing service fatigue
Fatigue crack growth failures. However, the DSTO’s and NLR’s experience is that only non-LEFM analyses are
Failure analysis
directly useful for determining service failure fatigue crack growth (FCG) behaviour. LEFM
analyses can assist in analysing FCG in full-scale, component and specimen tests, and these
analyses can be used – within limits – to predict some aspects of the FCG behaviour of in-
service structures and components. The DSTO has developed a hybrid method, the Effective
Block Approach (EBA) using both LEFM and non-LEFM analyses, from full-scale testing. The
EBA enables a wide range of FCG predictions for high performance aircraft.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction

The NLR and DSTO are the respective principal aerospace research organisations in the Netherlands and Australia. Both
have more than 40 years’ experience of failure analysis, covering a variety of aircraft types, numerous structures and com-
ponents, and both service and test failures. The present paper draws upon this experience to discuss the uses and limitations
of fracture mechanics, specifically Linear Elastic Fracture Mechanics (LEFM), in analysing the failures. The discussion is con-
fined to fatigue failures in metallic aircraft structures, since these types of failures predominate and are also the main con-
tributors to serious accidents and incidents [1–5]. This predominance of fatigue reflects the dynamic nature of aircraft
service loads and the relatively high design stress levels required to achieve lightweight structures.

2. Potential and actual uses of fracture mechanics in aircraft failure analysis

Fracture mechanics analyses may be considered for service failures, full-scale and component tests, and specimen/coupon
tests. For full-scale and component tests such analyses can be applied to both interim-detected cracking and cracks detected
during final teardown. An example of a full-scale test article is given in Fig. 1. The complex configuration of this component is
evident, and such complexity can and does pose problems for all types of analyses, as will be discussed in Section 5 of this
paper.
The types of analyses and their potential uses are specified here with the aid of Fig. 2:

⇑ Corresponding author.
E-mail address: Russell.Wanhill@nlr.nl (R.J.H. Wanhill).

1350-6307/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfailanal.2012.10.022
34 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

Fig. 1. Full-scale fatigue testing of three Boeing F/A-18 centre barrel bulkheads attached to the engine ducts. The bulkheads have highly complex
configurations.

Fig. 2. Schematic of fatigue crack growth from an initial discontinuity/crack size, ai, to the critical crack size, ac. The minimum reliably detectable crack size,
andi, is also indicated.

(1) Fatigue crack growth (FCG).


 Life to the maximum/critical crack size ac from an initial discontinuity/crack size ai.
 Life to the maximum/critical crack size ac from the minimum reliably detectable crack size andi.
 Feasibility of replacement/repair for remaining service or test life.
 Effect of aircraft usage severity (individual and fleet differences) on life.

The latter two uses require explanation. Failures or detected cracking require assessment of whether replacement or
repair of the failed or damaged components will enable the remaining service or test lives to be reached with or without
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 35

further replacements or repairs. FCG analyses based on service or test data can be used in these assessments. Similar analyses
may be used to assess the effects of aircraft usage, through its effects on service loads, on the in-service crack growth behav-
iour and life.

(2) Fracture toughness.


 Estimation of the maximum permissible crack size (not necessarily the critical crack size ac) for service and testing.

For example, the Royal Australian Air Force (RAAF) has a cracked structure residual strength (RS) requirement of
1.2  Design Limit Load (DLL) without failure [6]. The maximum permissible crack size aRS at the RS P 1.2DLL point is
not necessarily the critical crack size under service loading.

3. FCG analysis methods

In the present context there are basically three types of FCG analyses to consider:

 LEFM based analyses.


 non-LEFM based analyses.
 LEFM/non-LEFM based analyses: the Effective Block Approach (EBA).

3.1. LEFM based analyses

The starting point for these analyses is the fitting of Paris-type equations [7,8] to constant amplitude (CA) fatigue crack
growth data obtained from laboratory coupon tests in normal air. The basic equations are:
da  pffiffiffiffiffiffim
¼ C DK m ¼ C bDS pa ð1Þ
dN
pffiffiffiffiffiffi
where da/dN is the FCG rate, DK ¼ DS pa is the stress intensity factor range, C and m are empirical constants, and b is the
geometry correction factor. These equations describe a fit of the central part (Region II) of a generic (simplified) FCG curve,
see Fig. 3, obtained from testing in a normal air environment.
Provided b is constant or approximately constant, Eq. (1) can be directly integrated to give FCG lifetimes under CA loading,
as follows:
" #
2 1 1
For m – 2; Nf ¼  ð2Þ
ðm  2ÞCbm DSm pm=2 a0ðm2Þ=2 afðm2Þ=2

1 af
For m ¼ 2; Nf ¼ Ln ð3Þ
Cb2 ðDSÞ2 p a0

where a0 and af are any chosen initial and final crack sizes, respectively.
Eqs. (1)–(3), or similar ones, are often given in textbooks and sometimes used in research paper modelling, e.g. Carpinteri
and Paggi [9]. However, the geometry correction factor b usually varies for real structures and components, and integration

Fig. 3. Schematic da/dN versus DK diagram for long fatigue crack growth in normal air, showing three regions of fatigue crack growth. DKth is the threshold
stress intensity factor range; KIc and Kc are the plane strain and non-plane strain fracture toughnesses, respectively.
36 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

must then be done numerically. Furthermore, numerical integration of CA-based equations is generally necessary because
the load histories in service or during full-scale and component tests are usually variable amplitude (VA) and often complex.
Another complication (a minor one) is that Region II of the FCG curve can have multiple linear segments [10,11].
Owing to the foregoing complications, all LEFM based FCG prediction models estimate crack growth via a cycle-by-cycle
process using integration algorithms [12,13]. Fig. 4 surveys the types of models and some of those available. Detailed discus-
sions of the types of models are given in [13,14]. One of the most advanced ones, FASTRAN, is illustrated by the flow chart in
Fig. 5.

3.2. Non-LEFM based analyses

These analyses require Quantitative Fractography (QF) of fatigue fracture surfaces. This requirement should be regarded
as a merit rather than a hindrance, since QF crack growth data represent reality, whether they are from service failures or
full-scale, component and specimen tests. There are basically two kinds of QF measurements, depending on whether the load
history is nominally CA or VA:

(1) Nominally CA-FCG


 Measurements of fatigue striation spacings versus crack size, a, to obtain da/dN versus a plots.

Then derive a versus N plots numerically, or analytically by equation fitting. Examples of successfully used equations,
for Multiple Site fatigue Damage (MSD) in pressure cabin lap splices [15,16] and notched specimens simulating the
blade roots of a gas turbine disc [17], are:

da 1  Ba0 
¼ AeBa ! Nf ¼ e  eBaf ð4Þ
dN AB

where A and B are empirical constants.


(2) VA-FCG
 Measurements of frequent or infrequent crack growth progression markers (natural or imposed) versus crack size,
a, to obtain partial a versus N plots. Check for an analytical fit to ‘‘complete’’ the a versus N plots.
 Measurements of frequent marker spacings, and maybe striation spacings, versus crack size, a, to obtain da/dN ver-
sus a plots. Then derive a versus N plots numerically, or analytically by equation fitting.

The choice between these measurement methods depends on the marking characteristics; see e.g. Barter et al. [18,19].
Examples of analytical equations used for many service failures and test cracks in high performance aircraft [20,21]
are:

1 af
af ¼ a0 ekN ! Nf ¼ Ln ð5Þ
k a0

where k is an empirical constant that depends on the load spectrum and material. Eq. (5) represent exponential crack
growth, which is also the case for Eq. (3). However, Eq. (3) is for CA loading only and requires b to be constant, whereas
Eq. (5) do not.

Fig. 4. Overview of LEFM-based fatigue crack growth analysis models. After Iyyer et al. [12].
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 37

Fig. 5. Flow chart of FCG analysis using the strip yield/crack closure model FASTRAN [13]. The module at top left is a special one for calculating the crack
opening stress for short cracks.

3.3. LEFM/non-LEFM based analyses: Effective Block Approach (EBA)

The EBA is a hybrid of LEFM and non-LEFM analyses that evolved from extensive experience with full-scale testing of high
performance aircraft structures and components, combined with specimen (coupon) tests and QF of fatigue fracture surfaces
[22,23]. The EBA requires several conditions/assumptions and inputs:

 The load histories consist of relatively short sequences of blocks of VA spectrum loading, such that an FCG curve is pro-
duced from at least 30 load blocks [22].
 Each block of VA loading may be treated as if it were a CA cycle, and for each block load history the severity level is quan-
tified by an overall reference stress, Sref, usually the peak stress level.
 Specimen (coupon) tests at different Sref levels to supplement the results from full-scale tests, which are usually done at
one Sref level only.
 QF measurements of frequent marker spacings (natural or imposed) versus crack size, a, to obtain a versus NB plots, where
NB is the number of VA blocks.
 FCG rates can be modelled as power functions of the crack length.
 Selection of suitable FCG models and determination of their empirical constants from the full-scale and coupon test QF
data.

The models and their empirical constants are then used to predict FCG for different load history conditions pertaining to
different locations in the aircraft structures and components, and also to (minor) differences in service aircraft usage. The
predictions are described in more detail by Barter et al. [22] and Molent et al. [23]. Either LEFM or non-LEFM models are
used, depending on the quantified FCG behaviour. The basic equations are given here.

(1) LEFM (Paris-type) equations

da  pffiffiffiffiffiffimB
¼ C B bSref pa ð6Þ
dNB
If b is constant or approximately constant, Eq. (6) can be directly integrated to give FCG lifetimes, as follows:
" #
2 1 1
For mB – 2; NB;f ¼  ðm 2Þ=2 ð7Þ
ðmB  2ÞC B bmB Sm
ref p
B mB =2 ðm 2Þ=2
a0 B af B

1 af
For mB ¼ 2; NB;f ¼ Ln ð8Þ
C B b2 ðSref Þ2 p a0
38 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

As noted in Section 3.1, b usually varies for real structures and components, and integration of Eq. (6) must then be done
numerically.

(1) Non-LEFM equations (Frost and Dugdale exponential-type) [22–24]


 
For a versus N B plots; a ¼ a0 exp kB Sm mB
ref N B ! LnðaÞ ¼ kB Sref N B þ Lnða0 Þ
B
ð9Þ
 
da 1
For da=dNB versus Sref plots; Ln  ¼ LnðkB Sm
ref Þ þ mB LnðSref Þ
B
ð10Þ
dNB a
These (or similar) equations describe FCG for many high-strength structures and components tested under relatively
short sequence VA loading [20–23]. In particular, exponential FCG is often characteristic for ‘lead cracks’, i.e. the fastest-
growing cracks in particular locations [20,21].

4. Survey of DSTO and NLR FCG analyses

Table 1 surveys the types of FCG analyses for most of the well-investigated service and test examples from the DSTO and
NLR archives and publications. Note that there were no examples where exclusively LEFM based analyses were used. This is
discussed in Section 6 of the paper.
Tables 2 and 3 list the non-LEFM analysed structures and components. The EBA analysed component was an F/A-18 A/B
wing aft closure rib [23]. All the DSTO non-LEFM analyses used the VA-FCG approach and Eq. (5), see Section 3.2 of this pa-
per. Most of the NLR non-LEFM analyses were numerical, but two used the nominally CA-FCG approach, Eq. (4) and some
additional assumptions for analysing the FCG behaviour of pressure cabin longitudinal lap splices in several transport aircraft
[16,34,35].

5. FCG analysis method limitations and problems

5.1. Service and full-scale test FCG aspects

There are a number of actual and potential issues and problems to be considered and/or accounted for when conducting
FCG analyses for service aircraft and full-scale tests:

(1) Loads and stresses. The fatigue load histories are obviously important. In most cases the load history will be VA, but
sometimes nominally CA, e.g. Table 3. In both cases the stress ratios, R = Smin/Smax, should be known for computational
modelling. For VA load histories it is necessary also to know or determine the types of spectra, whether the load
sequences are short or long, and the frequency of occurrence and distribution (random or non-random) of peak loads
and underloads. This latter aspect is particularly important for QF measurements of FCG via crack growth progression
markers.
(2) Structures and components. These can have complex geometries, e.g. Fig. 1, resulting in stress concentrations and
eccentricities leading to secondary bending. The occurrence of secondary bending during service or simulated-service
loading is widespread in aircraft structures and is often unrecognized or poorly quantified. Notable examples are pres-
sure cabin longitudinal lap splices, which undergo both longitudinal and circumferential secondary bending that sig-
nificantly affect FCG from the stress concentrations provided by rivet holes [16,34,35].
Other problems are residual stresses in components and structural assembly (fit-up) stresses; and load-shedding,
owing to loss of stiffness, from cracked to uncracked components. These complications and their effects on FCG are
difficult or impossible to quantify prior to full-scale testing and QF measurements of FCG.
An essential aspect that greatly affects the viability and validity of FCG analyses is the range of initial discontinuity
sizes representing the effective pre-crack sizes (EPS) from which crack growth begins [20,21,39,40]. The initial discon-
tinuity sizes and EPS values are much smaller than the assumed initial crack sizes commonly used in FCG analyses, e.g.
the equivalent initial flaw sizes (EIFS) specified by the United States Air Force (USAF) damage tolerance requirements
[41].

Table 1
FCG analysis types used for 21 examples of fatigue cracking and failure.

FCG analysis type Failure category DSTO NLR


LEFM Service/test – –
Non-LEFM Service/test 4/6: Analytical 3/7: Numerical (8) and analytical (2)
EBA Test 1 –
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 39

Table 2
DSTO examples of non-LEFM based FCG analyses [20–31].

Aircraft/engine Component and material Service VA-FCG test


Aermacchi MB-326H Wing spar boom: AA 7075-T6 VA
Wing attachment boom: AISI 4340 steel 
Boeing F/A-18 A/B Trailing edge flap lugs: AA 7050-T7451 VA
Vertical tail root: AA 7050-T7451 
Centre barrel bulkhead: AA 7050-T7451 
Rudder hinge fitting: AA 7075-T73 
Horizontal tail spindle: AF1410 steel 
GD/LM F-111 Lower wing skin: 2024-T851 VA 
P&W 125B engine Bearing cage: AISI 4340 steel VA

(3) Miscellaneous. Environmental influences on service and full-scale test FCG behaviour are usually well-represented by
fatigue in normal air, as evidenced by the similar fractographic appearances and the absence of corrosion damage such
as pitting and oxide formation on the fracture surfaces. However, occasional examples of corrosion-induced damage
are observed for service failures [28,34,35]. The importance of an aggressive environment is that it accelerates FCG
compared to normal air, and this acceleration depends on the cycle frequency, see e.g. [42,43].
Fretting is a common initiator of fatigue [1], notably at faying surfaces and in fastener and pin-loaded holes [44]. Fret-
ting causes fatigue cracks to initiate very early in the life [37,45,46] and also increases early FCG rates [46,47].
Lastly, VA loading can change the fatigue fracture topography compared to CA loading [23,48]. Pronounced changes
can affect the FCG rates in unpredictable ways [48].

The issues and problems mentioned in (1) and (2) are of prime importance to the predictive capability of FCG models. The
‘miscellaneous’ aspects in (3) are usually (much) less important, which is just as well, since they are not explicitly considered
for modelling purposes.

5.2. LEFM based analyses

Fig. 5 and Table 4 show that LEFM FCG models have many requirements relating to the loads and stresses, point (1) in
Section 5.1. However, in the first instance these models take no account of point (2) except to use b-factors to cope with
geometry changes and stress concentrations. In particular, the models do not use actual short crack data. Instead they rely
on back-extrapolation of long crack growth data, obtained under CA loading, to describe short crack behaviour.
Two examples will be given here to illustrate the problems in using LEFM based analyses.

5.2.1. Example 1: helicopter simulated fuselage panels under VA loading


A ‘Helicopter Damage Tolerance Round Robin Challenge’ was begun in May 2002 and the results published in 2003
[49,50]. Participants were requested to predict the VA-FCG curves and lives, using LEFM based analyses, for two helicopter
simulated fuselage panels containing small starter flaws, see Fig. 6. The panels were machined from aluminium alloy AA
7010-T73651 plate. The participants were given the specimen geometry, Finite Element Method (FEM) stress intensity solu-
tions including b-factors, VA load history, and CA long crack growth data. Twelve organisations, from three continents, pro-
vided 27 ‘blind’ (uncalibrated by actual FCG data) predictions using eight different software analysis packages [49] and
options for load-interaction effects or non-interaction.
The predicted VA-FCG lives ranged from 240 flight hours to more than 12,500 flight hours, see Fig. 7. The test lives were
420 and 440 flight hours. Many of the predictions were within a factor of 2 of the test lives, which is usually considered
acceptable for blind predictions. However, some results were highly unconservative, even when the analyses used non-inter-
action options. Furthermore, none of the analyses gave the actual a versus N curve shapes [49]: accurate predictions of FCG
curve shapes are needed for establishing inspection intervals to enable timely detection of in-service cracking.
Overall, the results were unsatisfactory, though the participants must be given credit for making open-access blind pre-
dictions. Several reasons were given for the inadequacies [49,50]:

(1) Modelling curve fits. Software packages use curve fits to CA-FCG data. The curves often fit better in one region of crack
growth rates (see Fig. 3) and the fit quality can vary for different R values.
(2) Material test data. The predictions relied mainly on CA data in the near-threshold region. This region is characterised
by considerable data scatter, including DKth, and steep da/dN versus DK gradients that may be expected to strongly
influence the predictions.
(3) Stress intensity factors. The FEM solutions may have been in error, and other errors could have been made by the soft-
ware package representations of the solutions.

Some additional aspects have to be considered:


40 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

Table 3
NLR examples of non-LEFM based FCG analyses [4,15,16,31–38].

Aircraft/engine Component and material Service Test Analysis


Sikorsky S-61 Rotor blade: AA 6061-T6 CA Numerical
Northrop NF-5 Wing root lower skin: AA 7075-T651 VA Numerical
Fokker F100 Vertical stabilizer rib: AA 7175-T736 VA Numerical
Fokker F100 Fuselage lap splices: AA 2024-T3 CA Analytical
BAC 1-11 Fuselage lap splices: AA 2024-T3 CA Analytical
Airbus Mega Liner Barrel Window frame: AA 7175-T73 VA Numerical
Liner barrel Window panel: GLARE 3a VA Numerical
(A380 MLB) Passenger door beam: GLARE 3a CA Numerical
NH industries NH-90 Tail hinge beam: AA 7075-T7351 VA Numerical
P&W F100-PW-220 Air supply manifold support rod: In 718 VA Numerical
a
GLAss REinforced AA 2024-T3 laminates [37].

(4) Load history effects. The load spectrum was dominated by many low-amplitude high-R cycles interspersed with a few
underloads rather than peak loads [49]. This means that FCG load history effects would have been small. Also, the cho-
sen material would be expected to show only limited actual (and calculated) FCG load history effects, since it was a
high-strength AA 7000 series alloy [51,52].
(5) Panel configuration. The simulated fuselage panels were symmetrical on either side of the vertical axis, see Fig. 6, and
they were also one-piece machined from plate material. Hence they would not have been liable to significant second-
ary bending or load-shedding (load redistribution to the uncracked sides) except possibly during the later stages of
FCG.
(6) Starter flaw. The fatigue crack starter flaw had a 2 mm radius. This flaw size is well beyond realistic initial discontinuity
sizes [20,21,39,40] and actually in the long crack regime, i.e. crack sizes more than about 0.5 mm for aluminium alloys
[53,54].

All three of these aspects should have made it easier to obtain (reasonably) accurate predictions of the VA-FCG curves and
lives. This emphasizes the problems mentioned in points (1)–(3).

5.2.2. Example 2: simple specimens (coupons) under VA loading


The DSTO has done many VA-FCG tests on coupons of aluminium alloy AA 7050-T7451, whereby QF was used to deter-
mine the initial discontinuity sizes and FCG rates [18–23,27,40]. The results were compared with LEFM based analyses of the
crack growth behaviour. Fig. 8 gives a representative example, with predictions from the widely used software package AF-
GROW [55]. In this example the predictions have been calibrated by selection of a near-arbitrary initial crack depth of
0.05 mm to obtain optimum fits to the final FCG lives [23].
The fits are actually poor, especially for lower stress levels. The predicted FCG curve shapes are incorrect, such that early
FCG rates are severely under-predicted and the later FCG rates are over-predicted. Also, the curves can be unconservative, see
the set of blue-symbol data and the corresponding predicted curve. There may be several reasons for the poor results [22].
However, the main problem appears to be that LEFM based analyses use long crack CA-FCG data for the predictions and can-
not cope with actual short crack growth from realistic initial discontinuity sizes. For example, even small decreases in the
selected initial crack depth used for the predictions in Fig. 8 would have drastically increased the life predictions [23].

5.2.3. Summary
As mentioned at the beginning of Section 5.2, LEFM FCG analyses have many requirements relating to the loads and stres-
ses used in their predictions. This means there has to be a ‘trial and error’ selection of models and model parameters to ob-
tain the optimum analyses. Also, there are several actual and potential problems in using these models, notably the inability
to account for (unknown) secondary bending, load-shedding and residual stresses in actual structures and components, and
the application of long crack CA-FCG data to short crack growth predictions, including the use of unrealistic initial discon-
tinuity sizes.

Table 4
LEFM based FCG analysis model requirements.

 Basic (CA) FCG curves  Crack opening stress, Sop


– Stress ratios R; DK and DKeff – Long cracks
– Curve transitions – Short cracks, Stip
– Thresholds – Constraint factors, a, 2D and 3D
– Short cracks
 Load spectrum and load  Stress intensity factors, K
history
– Cycle counting, e.g. ‘‘Rainflow’’ – b: Geometry, stress concentrations
 Initial and final crack sizes  FCG cumulative damage
calculations
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 41

Fig. 6. Schematic diagram of a helicopter simulated fuselage panel (dimensions in mm). The heavy black arrow on Section XX points to the location of a
2 mm radius starter flaw. After Irving et al. [49].

In view of the foregoing remarks, and the examples given in Sections 5.2.1 and 5.2.2, it is evident that both uncalibrated
and calibrated LEFM based FCG life and curve predictions can be poor and unreliable, and that unconservative results are
possible. This conclusion agrees with previous results, especially with respect to uncalibrated LEFM models [12,56,57].

5.3. Non-LEFM based analyses

Table 5 lists the non-LEFM FCG analysis requirements. The most demanding requirement is QF measurements of FCG,
which are needed for both service and test failures. These measurements require considerable expertise and are labour-
intensive. Also, for service failures and some full-scale and component tests there is no guarantee that QF is possible: the
load history may be too complex and not provide natural markers. This is why there is much interest in providing markers
for tests [18,19].
If predictions are required, then supplemental FCG specimen (coupon) tests must be done using load spectra and load
histories representative of the known or expected service load histories, and employing natural or imposed markers
[18,19]. The usefulness of these tests may be limited or compromised by an inability to account for geometry changes, sec-
ondary bending, load-shedding and residual stresses in actual structures and components. The only way to compensate for
42 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

Fig. 7. VA–FCG life predictions and test lives for helicopter simulated fuselage panels. After Irving et al. [49].

these potential disadvantages is to thoroughly compare the full-scale and component QF results with those from coupons
and possibly more complicated specimens [33]. Several examples are given in Molent et al. [20].
It was mentioned in Section 4 that all the DSTO non-LEFM analyses used the VA-FCG approach and Eq. (5), while most of
the NLR non-LEFM analyses were numerical, with two using the nominally CA-FCG approach, Eq. (4). The choice of method
depended on the types of structures and components, the service and test load histories, and individual experience with suc-
cessful analyses. For example, Eq. (5), describing exponential-type FCG, have been successfully used by the DSTO for many
service and test cracks in high performance aircraft [20,21].

5.4. Effective Block Approach (EBA)

Table 6 summarises the EBA FCG analysis requirements, most of which are given in Section 3.3. The load spectrum, load
history and specimen test requirements stem from (a) the development of the EBA for analysing the FCG behaviour of high
performance aircraft structures and components, (b) the assumption that blocks of VA loading may be treated as CA cycles,
(c) the assumption that the severity level of different block load histories is quantified by a reference stress, Sref, and (d) the
need to supplement the results from full-scale tests, which are usually done at one Sref level only.
The most demanding requirement of the EBA is the large number of QF measurements. This is offset by the advantage of
an extensive data base that can be used to predict FCG for different load history conditions pertaining to different locations,
and also to minor differences in service aircraft usage, as mentioned in Section 3.3.
Like the non-LEFM analyses discussed in Section 5.3, EBA analyses may be partially compromised by an inability to ac-
count for geometry changes, secondary bending, load-shedding and residual stresses. However, this limitation applies only
to locations for which no QF data are available. Supplemental specimen tests, FEM analysis and strain gauge measurements
(if not already available from full-scale and component testing) may be needed to cope with these locations.

6. Selection and use of FCG analysis methods

6.1. LEFM based analyses

Uncalibrated (blind) LEFM based FCG analyses are unsuitable for service failures, since the predictions are too unreliable.
This situation is unlikely to change, owing to the many uncertainties and problems illustrated and discussed in Section 5.2.
Calibrated LEFM models fitted to service and test data, e.g. Fig. 8, may be used – within limits – for FCG life predictions to
indicate the effects of usage severity for service aircraft. These predictions should be limited to minor stress level changes
and differences in load histories:

(1) Stress levels. Interpolations are generally acceptable, but extrapolations should be restricted to stress level changes of
±10–20%. These limits are based on VA-FCG life data obtained for gust and manoeuvre spectrum loading [51,52].
(2) Load histories. Life predictions for minor load spectrum variations could be considered, notably for short sequence load
histories with frequent peak loads and underloads typical of high performance aircraft. However, there are no well-
defined guidelines, only suggestions and opinions based on altering the load histories to provide crack growth pro-
gression markers for QF [18,19,58].

For high performance aircraft the model calibrations and predictions would probably benefit from being included in EBA
analyses [22,23], which, however, require extensive data bases.
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 43

Fig. 8. Comparison of calibrated VA–FCG predictions and QF-determined VA–FCG for coupons of aluminium alloy AA 7050-T7451 tested with a tactical
aircraft spectrum. The characteristic stress is a measure of the VA load history severity, in this case the peak stress level. After Molent et al. [23].

Table 5
Non-LEFM based FCG analysis requirements.

 Load spectrum and load history  Numerical or preferably analytical fits to QF data
– Natural or imposed markers
 QF measurements of FCG  Predictions also require:
– CA loading: striation spacings – FCG specimen (coupon) tests
– VA loading: marker spacings – Initiala and final crack sizes
a
Derived from back-extrapolation of QF data to obtain the effective pre-crack sizes (EPS).

6.2. Non-LEFM based analyses

Non-LEFM based FCG analyses are the only ones that the DSTO and NLR have found to be directly applicable to service
failures, provided that QF measurements of FCG were possible. These analyses are also useful for full-scale, component
and specimen tests, again only if QF is possible.
Calibrated analyses derived from service and test data can be used for FCG life predictions, a versus N curves, and da/dn
versus a behaviour. Most usefully, these analyses can provide estimates of the inspectable lives and inspection intervals for
service aircraft, and repair or replacement options for failed or damaged components, i.e. whether the components can be
repaired or must be replaced in order to achieve the remaining service (or test) lives. Similar analyses may be used to assess
the effects of service usage severity via stress level interpolations of test data and limited extrapolation of test data for minor
variations in stress levels and load histories, see points (1) and (2) in Section 6.1.
For high performance aircraft the non-LEFM analysis of test results are known to benefit from being included in EBA anal-
yses [22,23].

6.3. Effective Block Approach (EBA)

The EBA is inherently restricted to the analysis of test data and therefore not directly applicable to service failures. How-
ever, the EBA’s flexibility in using an extensive data base and LEFM and non-LEFM analyses enables the prediction of FCG
behaviour for many locations in high performance service aircraft that experience broadly similar load histories. In turn,
these predictions enable estimates of the inspectable lives and inspection intervals; repair or replacement options for failed
or damaged components; and the effects of usage severity, subject to the limitations already mentioned in Sections 6.1 and
6.2.
44 R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45

Table 6
Effective Block Approach (EBA) FCG analysis requirements.

 Load spectrum and load history  FCG specimen (coupon) tests at different Sref levels
– Relatively short load blocks
– No high over- and underloads
– Natural or imposed markers  QF measurements of markers
 LEFM and non-LEFM fits to QF data depending on suitability  Initiala and final crack sizes
a
Derived from back-extrapolation of QF data to obtain the effective pre-crack sizes (EPSs).

7. Concluding remarks

The DSTO and NLR have more than 40 years experience of service and test fatigue failures from a variety of aircraft types,
structures and components. This experience has shown that only non-LEFM based fatigue crack growth (FCG) analyses that
rely on Quantitative Fractography (QF) data can be directly applied to service failures. LEFM based FCG analyses are too unre-
liable without being calibrated by the actual data, and this situation is unlikely to change.
Calibrated LEFM based analyses, using data from service, full-scale, component and specimen tests, may be used – within
limits – for FCG life predictions. However, the predictions can be poor and unreliable, especially for lower stress levels and
longer lives. Also, the predicted FCG curve shapes are incorrect. An important reason for this is that LEFM based analyses use
long crack CA-FCG data for predictions of short crack growth and cannot cope with actual short crack growth from realistic
initial discontinuity (crack) sizes.
A hybrid method, the Effective Block Approach (EBA), using both LEFM based and non-LEFM based analyses, has been
developed by the DSTO from full-scale testing of high performance aircraft structures, supplemented by specimen tests.
The EBA is flexible in the choice of analyses and enables a wide range of FCG predictions for service aircraft.

References

[1] Campbell GS, Lahey R. A survey of serious aircraft accidents involving fatigue fracture. Int J Fatigue 1984;6(1):25–30.
[2] Brooks CR, Choudhury A. Failure analysis of engineering materials. New York: McGraw-Hill; 2001.
[3] Findley SJ, Harrison ND. Why aircraft fail. Mater Today 2002;5(11):18–25.
[4] Wanhill RJH. Some notable aircraft service failures investigated by the national aerospace laboratory (NLR). Struct Integrity Life 2009;9(2):71–87.
[5] Tiffany CF, Gallagher JP and Babish IV CA. Threats to structural safety, including a compendium of selected structural accidents/incidents. USAF
Technical Report ASC-TR-2010-5002. Aeronautical Systems Center Engineering Directorate, Wright-Patterson Air Force Base OH (USA); 2010
[6] Royal Australian Air Force, Structural analysis, methodology – F/A-18A/B, ASI/2006/1114755 Pt1 (18), Issue 2, AL2, 7 December 2007.
[7] Paris PC, Gomez MP, Anderson WE. A rational analytic theory of fatigue. Trend Eng 1961;13:9–14.
[8] Paris PC, Erdogan F. A critical analysis of crack propagation laws. J Basic Eng, Trans ASME 1963;85D:528–34.
[9] Carpinteri A, Paggi M. A unified interpretation of the power laws in fatigue and the analytical correlations between cyclic properties of engineering
materials. Int J Fatigue 2009;31:1524–31.
[10] Wanhill RJH. Low stress intensity fatigue crack growth in 2024–T3 and T351. Eng Fract Mech 1988;30:233–60.
[11] Wanhill R, Barter S. Fatigue of beta processed and beta heat-treated titanium alloys. Dordrecht: Springer; 2012.
[12] Iyyer N, Sarkar S, Merrill R, Phan N. Aircraft life management using crack initiation and crack growth models – P-3C Aircraft experience. Int J Fatigue
2007;29:1584–607.
[13] Hu W, Tong YC, Walker KF, Mongru D, Amaratunga R and Jackson P. A review and assessment of current lifing methodologies and tools in Air Vehicles
Division. DSTO Research Report DSTO-RR-0321. DSTO Defence Science and Technology Organisation, Fishermans Bend, Victoria 3207, Australia; 2006.
[14] Schijve J. Fatigue of structures and materials. 2nd ed. Dordrecht: Springer; 2009.
[15] Eijkhout MT. Fractographic analysis of longitudinal fuselage lapjoint at stringer 42 of Fokker 100 full scale test article TA15 after 126350 simulated
flights. Fokker Report RT2160. Fokker Aircraft Ltd., Amsterdam (the Netherlands); 1994.
[16] Wanhill RJH and Hattenberg T. Fractography-based estimation of fatigue crack ‘‘initiation’’ and growth lives in aircraft components. NLR Technical
Publication NLR-TP-2006-184. Amsterdam (the Netherlands) National Aerospace Laboratory NLR; 2006.
[17] Wanhill RJH, Hattenberg T and Ten Hoeve HJ. Fractographic examination of three EGT Rim Specimens. NLR Contract Report NLR-CR-96451C. National
Aerospace Laboratory NLR, Amsterdam (the Netherlands); 1996.
[18] Barter SA and Wanhill RJH. Marker loads for quantitative fractography (QF) of fatigue in aerospace alloys. NLR Technical Report NLR-TR-2008-644.
National Aerospace Laboratory NLR, Amsterdam (the Netherlands); 2008.
[19] Barter SA, Molent L, Wanhill RJH. Marker loads for quantitative fractography of fatigue cracks in aerospace alloys. In: Bos MJ, editor. ICAF 2009,
bridging the gap between theory and operational practice. Dordrecht: Springer Science+Business Media; 2009. p. 15–54.
[20] Molent L, Barter SA and Wanhill RJH. The lead crack fatigue lifing framework. DSTO Research Report DSTO-RR-0353. DSTO Defence Science and
Technology Organisation, Fishermans Bend, Victoria 3207, Australia; 2010.
[21] Molent L, Barter SA, Wanhill RJH. The lead crack fatigue lifing framework. Int J Fatigue 2011;33:323–31.
[22] Barter S, McDonald M and Molent L. Fleet fatigue life interpretation from full-scale and coupon fatigue tests – a simplified approach. In: USAF
structural Integrity Program (ASIP) conference, 2005, Memphis, TN, USA.
[23] Molent L, McDonald M, Barter S, Jones R. Evaluation of spectrum fatigue crack growth using variable amplitude data. Int J Fatigue 2008;30:119–37.
[24] Frost NE, Dugdale DS. The propagation of fatigue cracks in test specimens. J Mech Phys Solids 1958;6:92–110.
[25] Clark G, Yost GS, Young GD. Recovery of the RAAF macchi MB326H – the tale of an ageing trainer fleet. In: Cook R, Poole P, editors. ICAF 97, vol.
I. Warley, West Midlands: Engineering Materials Advisory Services; 1997. p. 39–58.
[26] Athiniotis N, Barter SA, Van Blaricum T, Clark G, Jost G, Richmond M, et al. RAAF Macchi MB326H centre section lower boom fatigue testing. DSTO
Technical Report DSTO-TR-0328. DSTO Defence Science and Technology Organisation, Fishermans Bend, Victoria 3207, Australia; 1996.
[27] Barter S, Molent L, Goldsmith N, Jones R. An experimental evaluation of fatigue crack growth. Eng Fail Anal 2005;12:99–128.
[28] Barter S, Sharp PK, Clark G. The failure of an F/A-18 trailing edge flap hinge. Eng Fail Anal 1994;1:255–66.
[29] McDonald M and Molent L. Fatigue assessment of the F/A-18 horizontal stabilator spindle. DSTO Technical Report DSTO-TR-1620. DSTO Defence
Science and Technology Organisation, Fishermans Bend, Victoria 3207, Australia; 2004.
[30] Molent L, Barter SA. A comparison of crack growth behaviour in several full-scale airframe fatigue tests. Int J Fatigue 2007;29:1090–9.
R.J.H. Wanhill et al. / Engineering Failure Analysis 35 (2013) 33–45 45

[31] Barter SA, Lynch SP and Wanhill RJH. Failure analysis of metallic materials: a short course. NLR Technical Publication NLR-TP-2010-550. National
Aerospace Laboratory NLR, Amsterdam (the Netherlands); 2010.
[32] Wanhill RJH, De Graaf EAB and Van der Vet WJ. Investigation into the cause of an S-61N helicopter accident. Part I: Fractographic analysis and blade
material tests. NLR Technical Report NLR-TR-74103C. National Aerospace Laboratory NLR, Amsterdam (the Netherlands); 1974.
[33] Van der Linden HH. Modifications of flight-by-flight load sequences to provide for good fracture surface readability fatigue crack topography. In:
AGARD conference proceedings no. 376. Neuilly-sur-Seine, France: Advisory Group for Aerospace Research and Development; 1984. p. 11-1–11-22.
[34] Wanhill RJH, Hattenberg T and Van der Hoeven WA. A practical investigation of aircraft pressure cabin MSD fatigue and corrosion. NLR Contract Report
NLR-CR-2001-256. National Aerospace Laboratory NLR, Amsterdam (the Netherlands); 2001.
[35] Wanhill RJH, Koolloos MFJ. Fatigue and corrosion in aircraft pressure cabin lap splices. Int J Fatigue 2001;23:S337–47.
[36] Wanhill RJH. Fatigue of air supply manifold support rod in military jet engines. J Fail Anal Prev 2004;4:53–61.
[37] Wanhill RJH, Platenkamp DJ, Hattenberg T, Bosch AF, De Haan PH. GLARE teardowns from the MegaLiner Barrel (MLB) fatigue test. In: Bos MJ, editor.
ICAF 2009, bridging the gap between theory and operational practice. Dordrecht: Springer Science+Business Media; 2009. p. 145–67.
[38] Hattenberg T and Bos MJ. Fractographic investigation of the Lower Hinge Beam of the fatigue test article of the NH-90 tail module. NLR Contract Report
NLR-CR-2010-652. National Aerospace Laboratory NLR, Amsterdam (the Netherlands); 2010.
[39] Molent L. Fatigue crack growth from flaws in combat aircraft. Int J Fatigue 2010;32:639–49.
[40] Barter SA, Molent L, Wanhill RJH. Typical fatigue-initiating discontinuities in metallic aircraft structures. Int J Fatigue 2012;41:11–22.
[41] Military specification airplane damage tolerance requirements, MIL-A-83444 (USAF), 1974.
[42] Bradshaw FJ, Wheeler C. The influence of gaseous environment and fatigue frequency on the growth of fatigue cracks in some aluminum alloys. Int J
Fract Mech 1969;5:255–68.
[43] Wanhill RJH. Environmental effects on fatigue of aluminium and titanium alloys, Corrosion Fatigue of Aircraft Materials. AGARD Report No. 659. pp. 2-
1–2-37, Advisory Group for Aerospace Research and Development, Neuilly-sur-Seine, France; 1977.
[44] Forsyth PJE. Occurrence of fretting fatigue failures in practice. In: Waterhouse RB, editor. Chapter 4 in fretting fatigue. London: Applied Science
Publishers.; 1981. p. 99–125.
[45] Endo K. Practical observations of initiation and propagation of fretting fatigue cracks. In: Waterhouse RB, editor. Chapter 5 in fretting
fatigue. London: Applied Science Publishers.; 1981.
[46] Edwards PR. The application of fracture mechanics to predicting fretting fatigue. In: Waterhouse RB, editor. Chapter 3 in fretting
fatigue. London: Applied Science Publishers.; 1981. p. 67–97.
[47] Edwards PR, Ryman RJ and Cook R. Fracture mechanics prediction of fretting fatigue under constant amplitude loading. RAE Technical Report 77056.
Royal Aircraft Establishment, Farnborough, UK; 1977.
[48] Wanhill RJH, ‘t Hart WGJ, Schra L. Flight simulation and constant amplitude fatigue crack growth in aluminium-lithium sheet and plate. In: Kobayashi
A, editor. ICAF ‘91, aeronautical fatigue: key to safety and structural integrity. Tokyo, Ryoin Co., and Warley, West Midlands: Engineering Materials
Advisory Services; 1991. p. 393–430.
[49] Irving, PE, Lin, J and Bristow JW. Damage tolerance in helicopters: Report on the Round Robin challenge, American Helicopter Society 59th Annual
Forum. St. Louis, MIRA Digital Publishing; 2003.
[50] Vaughan RE and Chang JH. Life predictions for high cycle dynamic components using damage tolerance and small threshold cracks, American
Helicopter Society 59th Annual Forum. St. Louis, MIRA Digital Publishing; 2003.
[51] Wanhill RJH. Flight simulation fatigue crack growth testing of aluminium alloys: specific issues and guidelines. Int J Fatigue 1994;16:99–110.
[52] Wanhill RJH. Damage tolerance engineering property evaluations of aerospace aluminium alloys with emphasis on fatigue crack growth. NLR
Technical Publication NLR-TP-94177U. National Aerospace Laboratory NLR, Amsterdam (the Netherlands); 1994.
[53] Anstee RFW. An assessment of the importance of small crack growth to aircraft design, behaviour of short cracks in aircraft components. In: AGARD
conference proceedings no. 328. Neuilly-sur-Seine, France: Advisory Group for Aerospace Research and Development; 1983. p. 3-1–9.
[54] Anstee RFW and Edwards PR. A review of crack growth threshold and crack propagation rates at short crack lengths, Some Considerations on Short
Crack Growth Behaviour in Aircraft Structures. AGARD Report No. 696. Neuilly-sur-Seine, France: Advisory Group for Aerospace Research and
Development; 1983. p. 2-1–2-12.
[55] Harter JA. AFGROW user’s guide and technical manual. Air Vehicles Directorate Technical Report AFRL-VA-WP-TR-2004. Air Force Research Laboratory,
Wright-Patterson Air Force Base, OH 45433–7542, USA; 2004.
[56] Brot A and Matias C. An evaluation of several retardation models for crack growth prediction under spectrum loading. In: USAF Structural Integrity
Program (ASIP) Conference. GA, USA (Savannah); 2002.
[57] Sander M, Richard HA. Lifetime predictions for real loading situations–concepts and experimental results of fatigue crack growth. Int J Fatigue
2003;25:999–1005.
[58] De Jonge JB, Schijve J. Review of Session IV: general discussion, fatigue crack topography. In: AGARD conference proceedings no. 376. Neuilly-sur-Seine,
France: Advisory Group for Aerospace Research and Development; 1984. p. D-1–3.

Potrebbero piacerti anche