You are on page 1of 13

Engineering Failure Analysis 11 (2004) 547559

www.elsevier.com/locate/engfailanal

Estimating fatigue life under variable amplitude


loading through quantitative fractography A case study
J.R. Tarpania,*, C.O.F.T. Ruckerta, M.T. Milana, R.V. Silvaa, A. Rosato Jr.b,
R.N. Pereirab, W.W. Bosea, D. Spinellia
a

Materials, Aeronautic and Automotive Engineering Department, Engineering School of Sao Carlos, University of Sao Paulo,
Sao Carlos-SP, 13.566-590, Brazil
b
Embraer S/A, Sao Jose dos Campos-SP, 12.227-901, Brazil
Received 10 September 2003; accepted 27 September 2003

Abstract
Quantitative fractography techniques have been implemented and frequently used in the Failure Analysis Laboratory
LANAF, at the Engineering School of Sao Carlos, for fatigue life estimations of structural components, whose
fracture surfaces are subjected to detailed inspection with the aid of scanning electron microscopes. This work describes
one of the recent activities in progress at the LANAF, in which fatigue crack initiation and propagation lives have been
estimated for an idealized aeronautical part tested in the laboratory, under variable amplitude loading condition
VAL, i.e., ight simulation testing. Fractographic reconstitution of sub-critical crack growth has been performed
through the identication of marking load patterns left in the wake of the propagating crack, which have been correlated to the signicant load levels applied during the fatigue test. A semi-automated procedure to estimate fatigue lives
under VAL has been developed and implemented by which a virtual marking load pattern is generated and compared
to the real pattern determined fractographically.
# 2004 Elsevier Ltd. All rights reserved.
Keywords: Aluminium alloys; Fatigue markings; Fractography; Marker loads; Quantitative fractography

1. Introduction
Fractographic techniques using scanning electron microscopes (SEM) are extensively used in failure
analysis of aeronautical components and structures [15]. The more advanced procedure employed on the
fatigue life estimation of in-service or in-test fractured components undergoing variable amplitude loading
VAL, is depicted in Fig. 1. The technique consists basically of generating a virtual pattern of fatigue
load marks provided by the applied loading spectrum, using preferably computational techniques, and
compares it with the pattern actually recorded on the fracture surface of the component, obtained through
microscopic survey. In summary, to be successful this procedure depends basically on the correct
* Corresponding author. Tel.: +55-16-273-9590; fax: +55-16-273-9574.
E-mail address: jrpan@sc.usp.br (J.R. Tarpani).
1350-6307/$ - see front matter # 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfailanal.2003.09.004

548

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Fig. 1. Schematic of the failure analysis procedure using quantitative fractography [4].

interpretation of the surface marking pattern printed by the higher cyclic loads (i.e. marker loads)
experienced by a ying aircraft or a tested structural component.
Fig. 2 shows typical load marking patterns impressed by the action of critical load sequences on the
fracture surface of an aeronautical component fatigue tested in the laboratory.
In order to estimate fatigue lives spent in both crack initiation and propagation stages, it is necessary to
obtain a plot similar to Fig. 3. From the numerical integration of the area under this curve, one can also
determine the number of ights as a function of the crack size, as well as the crack growth rate in the
structure or component analyzed.
The correct determination of fatigue crack initiation and propagation lives has obvious and important
implications for several steps of aircraft fabrication and maintenance, such as materials selection, component design, manufacturing processes, non-destructive inspection programs, and the ultimate failure
analysis as well.
In this work, an idealized aeronautical component was fatigue tested in a laboratory environment
in order to simulate its in-service performance under VAL. After testing, the fracture surfaces were
thoroughly analyzed in a SEM. The purpose was to determine the resulting crack surface marking
patterns, which were compared to those derived from a computer program, which had the simulated
ight load history as the input. The methodology described in Fig. 1 has been strictly followed, aimed at
determining both crack initiation and propagation lives of the fatigue-tested component.

2. Materials and component tested


A 1.5 in.-thick rolled plate of a 7475-T351-aluminium alloy, specication SAE-AMS4204 [6] was
investigated in this study. Table 1 presents the chemical composition (wt%) of the alloy as determined via
optical emission spectroscopy. Table 1 also supplies the allowable quantities for the elements according to
the SAE specication, which are strictly obeyed by the Al-alloy tested.
The microstructure of the aluminium alloy is showed in Fig. 4. Texture is found in the rolling
direction, L. Wide non-recrystallized regions and segregation can be observed in LS, and particularly,
in TS planes.

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

549

Fig. 2. Typical load marking patterns (arrowed) developed on the fracture surface of a fatigue-tested component: (a) near crack
origin; (b) intermediate location; (c) nal fracture position.

3. Experimental and analytical


An idealized aeronautical component designed by LANAF was fatigue tested under a VAL spectrum
expected to occur in the lower wing skin panels of commercial aircraft. Fig. 5(a) presents the main fractured
ligament F2 ! F1, which has been analyzed in this study. The black arrow in Fig. 5(a) indicates the
location of the fatigue crack site, i.e. a corner crack.
The cyclic loading conguration applied to the part is given in Fig. 5(b). A bending moment is
introduced in the tested component, whose magnitude and prole depend on the fasteners array employed
to connect it up to the auxiliary plates, which in turn are gripped by the servo-hydraulic closed-loop test
machine. Therefore, tensile stresses are induced in the surface of the sample that is in close contact with the
thicker plates. Consequently, a compression loading is generated in the opposite face of the tested part.

550

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Fig. 3. Fatigue crack growth data as determined by quantitative fractography [2].


Table 1
Chemical composition (wt%) of the aluminium alloy tested in this study
Element

Weight (%)

Si
Fe
Cu
Mn
Mg
Cr
Ni
Zn
Ti
Ca
P
Pb
Sb
Sn
Sr
V
Zr
B
Cd
Co
Al
Impurities

0.051 (max 0.10)


0.079 (max. 0.12)
1.67 (1.21.9)
0.011 (max 0.06)
2.265 (1.92.6)
0.253 (0.180.25)
0.004 (max 0.05)
5.909 (5.26.2)
0.020 (max 0.06)
ND (max 0.05)
0.010 (max 0.05)
0.002 (max 0.05)
0.004 (max 0.05)
0.001 (max 0.05)
ND (max 0.05)
0.002 (max 0.05)
ND (max 0.05)
0.0008 (max 0.05)
ND (max 0.05)
0.002 (max 0.05)
Balance
0.026 (max 0.15)

Limiting values according to SAE-AMS4204 are provided in parenthesis. ND: non-detected.

Fig. 6 shows two ight block sequences imposed on the component during the laboratory test. In total,
20 blocks of 6000 ights were sequentially applied to the component, therefore performing 120,000
simulated ights.

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

551

Fig. 4. Tri-dimensional views of the microstructure of the aluminium alloy: (a) lower and (b) higher magnitude.

Fig. 5. (a) Partial view of the VAL-fatigue-tested component. The black arrow points out the approximate location of the fatigue
crack site. The full length of the failed ligament F2 ! F1 is 23 mm; (b) schematic of the cyclic loading arrangement imposed on the
tested component.

552

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Fig. 6. Two loading blocks applied sub-sequentially to the component tested, with the severest ight, numbered 1, occurring once at
each applied block of 6000 ights.

After testing, the components fracture surfaces were inspected in a SEM operating at a voltage of 20 kV
in both secondary and backscattered electrons imaging modes. The chief interest was to elucidate the entire
load marking nuances left on the wake of the propagating crack. This task was accomplished by using
image magnications ranging from 10 to 20,000.
A computer program was specically developed to generate virtual marker load patterns from the
available ight load history, which were compared to marking patterns determined fractographically.
During the programming, special attention was devoted to VAL-variables such as maximum tensile stress
peaks and eective number of tensile stress cycles within a block of simulated ights, since reliable
criteria had to be developed concerning the width of virtual marking bands and band spacing as well. Once
the software provided the virtual pattern of load marking bands expected in one block of ights, it was
straightforwardly compared to that real pattern, in an operation that resembled the reading of bar codes.
The procedure was repeated until a minimum matching between virtual and real patterns was achieved.
When a successful correlation was reached, the ight blocks counting along the fracture surface could be
accomplished manually.

4. Results and discussion


4.1. Metallurgical examination of the crack nucleation site
Fig. 7 exhibits the microstructure of the alloy in the fatigue crack initiation site region, which has been
indicated by an arrow in Fig. 5(a). The arrow in Fig. 7 points out distinct neighbouring second-phases,
respectively black and white colours, located near the fatigue crack nucleation site. The particles were
analyzed and identied by energy dispersive X-ray (EDX) analysis. These metallurgical stress concentrators near the surface of the fatigued component suggested that microstructural heterogeneities could
have led to the failure of the component in the ight simulation test.

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

553

Fig. 7. Microstructural aspects of near fatigue crack site region, where the arrow points out two distinct second phases. The crack site
location, the borehole position and the fatigue crack surface are identied.

Fig. 8. EDX spectrum of (a) white and (b) black phases located in the neighbourhood of the fatigue crack nucleation site in the
borehole F2.

554

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Table 2
Chemical composition (wt%) of the intermetallic phases shown in Fig. 7
Element

White phase (wt%)

Black phase (wt%)

O
Al
Mn
Mg
Cu
Zn

7.03
23.65
0.33
ND
67.76
1.24

0.43
89.47
ND
1.13
1.95
7.03

The copper and zinc contents are highlighted for the white and black phases, respectively. ND: non-detected.

Fig. 9. (a) Top view of the fatigue crack nucleation site pointed out by a white arrow; (b) contact marks due to the friction of the
threaded fastener with the internal surface of borehole F2, indicated by black arrows; (c, d) detail of the marks, where white arrows
indicate the marks and the black one the crack site position; (e) cross-sectional view of one surface mark generated during the
disassembly of the jig, where the white arrow indicates the direction of the movement for the fastener removal.

4.2. Microchemical analysis of the phases


Figs. 8(a) and (b) show EDX data for the black and white colours phases depicted in Fig. 7. Table 2
presents their chemical composition, identied as Cu-rich and Zn-rich phases, respectively.

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

555

4.3. Contact marks on the borehole F2


Fig. 9(a) displays a partial top view of the fracture surface of the component, in which the arrow points
out the site of the fatigue crack initiation. The dark stain exhibited in this region occurred due to surface
contamination by accidental contact with lubricant oil drained from the mechanical test system.
The black arrows in Fig. 9(b) indicate some marks left by the contact of the fasteners thread screw with
the internal surface of the hole identied as F2 in Fig. 5(a). Notice that one of these contact marks lined up
exactly with the fatigue crack nucleation site, which is pointed out by the white arrow. However, none of
the marks reached the fatigued fracture plane, Fig. 9(c) and (d), in which the crack site is indicated by a
black arrow in Fig. 9(d). In fact, a more careful analysis showed that the marks left by the contact between
the fastener thread and the borehole F2 were indeed generated during the disassembly of the jig to carry
out the procedures of failure analysis. This is clearly veried in Fig. 9(e), which corresponds to a crosssectional view of the part. Here, the white arrow points out the fasteners withdrawal direction, where one
can see the material pulled out from the borehole due to friction with the threaded screw.
Fig. 10(a) shows a detailed view of the fatigue crack site of the fractured ligament F2 ! F1 (see also
Fig. 5a), where a non-metallic inclusion was located, as indicated by an arrow. This particle was identied
by EDX microanalysis as a calcium-sulphur rich phase. The corresponding energy absorption spectrum is
provided in Fig. 10(b).
Fig. 11 presents three SEM-microfractographs taken from a intermediate position of the fracture path
F2 ! F1, at distinct image magnications. The loading marks printed on the fracture surfaces are numbered in accordance with the ight codes established in Fig. 6, i.e. severest ight sequence which originated
this marking pattern. Notice the perfect correspondence between the fractographic marking sequence and
the ight sequence applied to the structural component, the latter depicted in Fig. 6. An extra marking, i.e.
not generated by the action of a severe ight, is identied in Fig. 11(b). This mark was probably due to an
overload experienced by the tested component in the automatic re-start of the mechanical testing system
just after the electric energy power supply had been re-established.
Fig. 12 shows an overall view of the cracked ligament F2 ! F1, in which the white arrows indicate the
sequential positions of occurrence of the severest ight 1. The conclusion, based on this gure is that at
least eight full blocks of 6000 ights have been applied to extend the fatigue crack along the whole ligament

Fig. 10. (a) Non-metallic inclusion located at the crack site of the fatigue-fractured ligament F2 ! F1; (b) energy absorption
spectrum resulting from EDX microanalysis.

556

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Fig. 11. SEM-microfractographs taken from the fracture surface of the fatigue-tested component at three dierent magnitudes of
image: (a) 70, (b) 230, and (c) 700. Arrows indicate the microscopic or local direction of crack growth along the fractured
ligament F2 ! F1.

assessed. Still in Fig. 12, the black arrow points out the fatigue crack nucleus position, which was rst seen
in Figs. 5(a), 7, 9(a), (b), (d) and 10(a).
Fig. 13 presents the resulting plot obtained from the fractographic analysis performed in this study,
according to the graphic format displayed earlier in Fig. 3. Two distinct methods have been used in order
to correlate the experimental data points, namely, general power-law and point-to-point linear tting.
Once the set of data points presented in Fig. 13 had been tted, the numerical integration of the area
under the tting curves was performed.
Fig. 14 plots results from both processes of numerical integration, by which one can estimate fatigue
crack initiation and propagation lives of the aeronautical part tested in fatigue. It can be observed that
power-law data tting predicts longer fatigue crack growth periods than a linear t, and, consequently,

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

557

Fig. 12. Sequential positions occupied by the severest ight number 1 along the fatigue crack path on the ligament F2 ! F1.

Fig. 13. Fatigue crack growth data determined fractographically from the tested component.

Fig. 14. Fatigue crack growth estimations according to distinct tting methods: (a) number of ights vs. crack depth; (b) vice-versa.

558

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

Fig. 15. Mean crack growth rate estimations according to the equivalent ight concept: (a) plotted against crack depth; (b) plotted
against number of simulated ights.

shorter crack nucleation periods, which is inferred to be 49,500 ights according to the former method
(recalling that total life of the component is 120,000 ights). On the other hand, according to the point-topoint linear data tting method, about 60,000 ights were necessary to start the crack growing in fatigue.
Therefore, from Fig. 14, two limiting situations can be postulated regarding periodic non-destructive
inspection and failure prevention of similar in-service components: rst, if fatigue crack initiation is the
main concern, i.e. fail-safe approach, the most conservative method is the power-law integration; second
and conversely, if cracking has already been established in the structural component, propagation becomes
of prime interest, i.e. damage-tolerant approach, and linear point-to-point integration is the most
conservative method.
Fig. 15 plots fatigue crack growth rate in function of crack depth and number of ights, respectively,
considering both integration methods shown in Fig. 13.
Almost imperceptible dierences are established in Fig. 15(a), regarding initial stages of crack growth.
Major dierences are noticed at higher levels of crack extension only. On the other hand, Fig. 15(b) allows
one to clearly dierentiate linear vs. power-law methods, as a result of the large dierences between fatigue
crack growth estimations exhibited in Fig. 14.

5. Concluding remarks
This article described one of the recent activities conducted in the Failure Analysis Laboratory
(LANAF) of the Engineering School of Sao Carlos in the eld of quantitative fractography.
In this study, crack initiation and propagation lives have been estimated for a fatigue tested aeronautical
part subjected to VAL, which has been imposed under laboratory controlled conditions, i.e. ight
simulation testing.
Fractographic reconstitution of sub-critical crack growth has been performed through the identication
of marking load patterns left in the wake of the propagating crack, which have been correlated to
signicant load levels applied during the fatigue test.

J.R. Tarpani et al. / Engineering Failure Analysis 11 (2004) 547559

559

A semi-automated procedure to estimate fatigue life under VAL has been developed and implemented,
by which a virtual marking load pattern is generated and straightforwardly compared to the real one
determined fractographically.
Implications for the adoption of more or less conservative life estimations methods in the structural integrity
program of commercial eets have been discussed on the basis of both fail-safe and damage-tolerant
approaches.

Acknowledgements
The authors wish to express their gratitude to Embraer S/A for supplying the material tested and for
valuable discussions.

References
[1] Abelkis PR. Use of microfractography in the study of fatigue crack propagation under spectrum loading. Fractography in
Failure Analysis, ASTM STP (Special Technical Publication) 1978;645:21334.
[2] Wiebe W, Dainty RV. Fractographic determination of fatigue crack growth rates in aircraft components. Can Aeronaut Space J
1983;27(2):10717.
[3] Dainty RV. Use of marker blocks as an aid in quantitative fractography in full-scale aircraft fatigue testing: a case study.
Fractogr Ceramics Metal Failures, ASTM STP 1984;827:285308.
[4] Goldsmith NT, Clark G. Analysis and interpretation of aircraft component defects using quantitative fractography. Quantat
Meth Fractogr, ASTM STP 1990;1085:5268.
[5] Goldsmith NT, Clark G, Barter SA. A growth model for catastrophic cracking in an RAAF aircraft. Eng Failure Anal 1995;3(3):
191201.
[6] Anon. SAE Aerospace Material Specication for Aluminium Alloy Plate 7475-T7351, Designation AMS 4202C, 1989.

You might also like