Sunteți pe pagina 1din 18

Available online at www.sciencedirect.

com

Engineering Fracture Mechanics 75 (2008) 3226–3243


www.elsevier.com/locate/engfracmech

A load history generation approach for full-scale


accelerated fatigue tests
J.J. Xiong a, R.A. Shenoi b,*
a
Aircraft Department, Beihang University, Beijing 100083, People’s Republic of China
b
School of Engineering Sciences, University of Southampton, Southampton SO17 1BJ, UK

Received 20 August 2007; received in revised form 23 October 2007; accepted 3 December 2007
Available online 23 December 2007

Abstract

This paper seeks to establish a load history generation approach for full-scale accelerated fatigue tests. Primary focus is
placed on the load cycle identification such as to minimize experimental time while having no significant effects on the new
generated load history. The load cycles extracted from an original load history are identified into three kinds of cycles
namely main, secondary and carrier cycles. Then the principles are presented to generate the load spectrum for accelerated
tests, or a large percentage of small amplitude carrier cycles are deleted, a certain number of secondary cycles are merged,
and the main cycle and the sequence between main and secondary cycles are maintained. The core of the generation
approach is that explicit criteria for load cycle identification are established and equivalent damage calculation formulae
are presented. These quantify the damage for accelerated fatigue tests. Three validation examples of its application for the
generation approach of accelerated load histories are given in the paper. Good agreement of experimental lives between the
original and generated load histories is obtained. Finally, the generation approach of accelerated load histories is applied
to the full-scale accelerated fatigue test of helicopter tail, demonstrating the practical and effective use of the proposed
approach.
Ó 2007 Elsevier Ltd. All rights reserved.

Keywords: Load cycle; Load history; Equivalent damage; Rainflow count; Original load history; Generated load history; Main cycle;
Secondary cycle; Carrier cycle

1. Introduction

In practice, the Palmgren–Miner linear accumulation damage rule [1,2], the S–N curve representing the
material performance determined from constant-amplitude tests, and load cycles defined using the rainflow
algorithm [3] are often used for fatigue life prediction of a structure subjected to a random load. Although
in most cases this is the best available method, the accuracy of the approach is often quite low. Consequently,
the method is generally used in the design stage, when the accuracy of fatigue life predictions is less important

*
Corresponding author. Tel.: +44 23 8059 2316; fax: +44 23 8059 3299.
E-mail address: r.a.shenoi@ship.soton.ac.uk (R.A. Shenoi).

0013-7944/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engfracmech.2007.12.004
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3227

Nomenclature

a0 initial crack length


acr critical crack length related to the stress level
C undetermined constant of fatigue crack growth rate da/dN–DK curve
D(Sa, Sm) damage resulted from a stress cycle of (Sa, Sm)
K stress intensity factor
Kc fracture toughness
Km mean value of stress intensity factor
m1 exponent of fatigue da/dN  DK formula
m2 exponent of fatigue da/dN  DK formula
N fatigue life
N(Sa, Sm) fatigue life with regard to fatigue stress level (Sa, Sm)
Q transformation variable
R stress ratio
Sa stress amplitude
Sm mean stress
Smax stress peak
Smin stress valley
a material constant
b(a) correction coefficient of stress intensity factor for the plate with finite width
r stress
rB tensile ultimate strength
rs yield stress
m Poisson ratio
DK stress intensity factor range
DKth fatigue crack growth threshold value

and when the experiments can be relatively cheap. For final evaluation of fatigue life, fatigue tests on compo-
nents or whole constructions under variable amplitude loading are performed extensively [4]. All of this testing
involves generating field or service loading using lengthy complex variable amplitude histories. These histories
may be generated by means of one of the following two methodologies.

(1) Using standardised spectra such as the FALSTAFF [5,6] or TWIST [7,8], WASH I [9], load–time his-
tories in the aircraft industry (in the design phase) are generated on the basis of a mission analysis
[10–13]. However, the histories relative to the missions are obtained by instrumenting components
and vehicles, which are then operated under varying service conditions. As is well known [14–18], the
actual load–time histories often contain a large percentage of small amplitude cycles where the fatigue
damage associated with these small amplitude cycles can be small. As a result, in many cases, small
amplitude cycles are deleted from these histories in order to produce representative and meaningful
yet economical testing. Schijve et al. [14] investigated the removal of small load ranges from the manoeu-
vre-dominated FALSTAFF load sequence. The deleting methodologies are either stress based [15,16] or
strain based [17,18]. These techniques attempt to remove cycles with negligible changes to the fatigue
damage or try to quantify a percentage change in the damage due to their deleting. However, deleting
of variable amplitude histories can drastically affect both crack initiation life and crack growth life
and this is highly dependent upon material, history, load level and sequence [18,19]. When generating
a load history it is essential that besides parameters such as amplitude, mean and the number of load
cycles, the sequence of the load cycles is taken into account. This sequence can often substantially influ-
ence a fatigue crack’s growth rate.
3228 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

(2) A new load history is generated on the basis of a distribution of load cycles extracted from the original
load history. The methods range from classical autoregressive methods [20,21] which are used to gener-
ate uncompressed load histories, to Markov methods [22,23] which are used to generate compressed load
histories. A common weakness of these methods is that the generation of new load histories is not based
on the rainflow counting method, the method most often used for extracting load cycles when an esti-
mation of the correlation between dynamic loads and the structure’s fatigue life is performed. An advan-
tage of the rainflow method is not only that the load cycles extracted from a complex, random sequence
with this method correspond to the closed hysteresis loops in the diagram, but it also that it enables the
sequence of load cycles to be taken into consideration when calculating the fatigue damage. Rychlik and
co-workers [23,24], for example, suggest that a rainflow matrix, which represents the distribution of load
cycles, should first be transformed into a corresponding Markov matrix; the latter is then used for the
generation of the new load history. With this approach the new load history is composed directly from
load cycles. However, these methods have two weaknesses. Firstly, the generated load history can be
composed only of those load cycles that were extracted from the original load history and secondly,
the information about the sequence of load cycles is lost.

The problem is that none of the above-mentioned methods alone fulfils all the requirements that should be
(at least formally) considered when generating new random load histories. These requirements are as follows:

 load cycles, extracted from original history, should correspond to the rainflow algorithm, because rainflow
cycle counting is a standard procedure for determining damage events in variable amplitude loadings;
 the information about the sequence of load cycles should be maintained;
 the damage resulting from the new generating load history should be the same as that from the original load
history;
 the new generated load histories should be of shorter length than the original load history to decrease the
test time, through deleting the small load cycles and merging the smaller ones.

Hence, it is desirable to develop a new method for the generation of load histories, which will pay regard to
all of these requirements. This work seeks to establish a load history generation approach for full-scale accel-
erated fatigue tests through deleting small amplitude carrier cycles and merging a certain number of secondary
cycles. Primary focus is placed on load cycle identification and equivalent damage calculation.

2. Load history generation principle

An original load history (shown in Fig. 1) always consists of the representative load cycles (shown in Fig. 2a
and b), which are termed as the compressed load histories [24,25]. Using the rainflow counting method, as sug-
gested by Amzallag et al. [24], the load cycles can be extracted from compressed load histories. The rainflow
counting method can be chosen to extract the remaining load cycles from a residuum that remains after count-
ing the load history [25]. By means of the rainflow counting method a load cycle is extracted from the

Fig. 1. An original load history.


J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3229

Fig. 2. Representative load cycle: (a) hanging load cycle; (b) standing load cycle.

compressed load history if four consecutive points A, B, C and D in the compressed load histories fulfil the
following relations (see also Fig. 2a and b):
jxC  xB j 6 jxB  xA j ð1Þ
jxC  xB j 6 jxD  xC j ð2Þ
The amplitude and the mean of the load cycle that is extracted with the rainflow method then are
jxC  xB j
Sa ¼ ð3Þ
2
xC þ xB
Sm ¼ ð4Þ
2
with the reversal point C being a half-point of the load cycle. In this way the load cycle is completely defined
with a pair of reversal points (B, C) from the compressed load histories. The relative position of these two
reversal points defines an orientation of the load cycle. If the reversal point B is higher than the reversal point
C, then the reversal point C corresponds to a valley and the load cycle is included in the examined history as a
hanging load cycle in the growing branch of a larger load cycle. If the reversal point B is lower than the rever-
sal point C, then the reversal point C corresponds to a peak and the load cycle is included in the examined
history as a standing load cycle in the falling branch of a larger load cycle shown in Fig. 2b. It is obvious that
all of the load cycles can be extracted from an original load history using the rainflow count method men-
tioned above. These extracted load cycles can be classified into three kinds of load cycles, namely main, sec-
ondary and carrier cycles.
Main cycles are the few larger load cycles, which probably cause significant damage to a structure; thus the
main cycle should be maintained during the new load history generation. An original history often contains a
large percentage of small amplitude cycles and the fatigue damage associated with these small amplitude cycles
can be small. These small amplitude and high frequency load cycles are termed as carrier cycles. As a result, in

Fig. 3. Deleted carrier cycle of hanging load cycle from original load history.
3230 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

many cases, carrier cycles are deleted from the original history in order to produce a representative and mean-
ingful yet economical testing load history (shown in Fig. 3). From Fig. 3, it is shown that the principle is pro-
posed to delete a carrier cycle as follows:

(a) Based on the rainflow counting, two adjacent and sequential secondary cycles (A, D) and (B, C) can be
extracted from an original load history A–B–C–D.
(b) According to the criterion of deleting small loads, the carrier load cycle (B, C) can be deleted; the remain-
der load cycle is (A, D). This remainder cycle (A, D) is used to replace the two original secondary cycles
(A, D) and (B, C).
(c) The new load history A–D is used to replace the original load history A–B–C–D.

As mentioned above, it is possible to define a secondary cycle as one larger than a carrier cycle but smaller
than a main cycle. Because of the great numbers of secondary cycles in any original load history, adjacent and
sequential secondary cycles should be merged into a new secondary cycle (shown in Fig. 4). From Fig. 4, one
can deduce a procedure to merge the adjacent and sequential secondary cycles as below.

(a) Based on the rainflow counting procedure, two adjacent and sequential secondary cycles (A, D) and
(B, C) can be extracted from an original load history A–B–C–D.
(b) According to the equivalent damage formulations given below, in case of the minimum stress of the new
secondary cycle being equal to or less than one of the two adjacent and sequential secondary cycles, a
new secondary cycle (A, D0 ) can be determined from above two secondary cycles (A, D) and (B, C). This
new secondary cycle (A, D0 ) is used to replace the two original secondary cycles (A, D) and (B, C); the
merging of two adjacent and sequential secondary cycles (A, D) and (B, C) is thus accomplished.
(c) The new load history A–D0 is used to replace the original load history A–B–C–D.

It is worth noting that if the new merged secondary cycle becomes a main cycle then it is necessary to stop
this merging. Thus information about the sequence of load cycles is maintained in order to maintain a similar
interaction effect. Further, the new generated load history is shorter in length than the original load history.
The generation process of a new load history can be represented schematically by a block diagram as shown in
Fig. 5.
Fatigue cracks generally appear at the surface of the material or at large inclusions, resulting from high
stresses, surface roughness, fretting, corrosion, etc. Fatigue crack growth on a macroscopic level usually
occurs perpendicular to the main principal stress and is dependent on the material, the material thickness
and the orientation of the crack relative to principal material directions. Furthermore, the crack growth
depends on the cyclic stress amplitude, the mean stress and the environment. The crack growth rate, denoted
da/dN, has become an important ‘‘material property” to characterize fatigue crack propagation for constant
amplitude loading. Generally, the crack growth rate is presented as a function of the stress intensity factor
range DK for different stress ratios R, material thicknesses, and different environments. Various deterministic
fatigue crack growth rate functions have been proposed in the literature. The functions can be represented by a
general form [26,27]:

Fig. 4. A new secondary cycle merged from two adjacent and sequential secondary cycles.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3231

Fig. 5. A block diagram of the load history generation for accelerated tests.

daðtÞ
¼ F ðDK; K max ; R; S; aÞ ð5Þ
dt
where da/dt is crack growth per flight time or cycle and F(DK, Kmax, R, S, a) is a non-negative function. Some
crack growth rate functions, such as Paris–Erdogan model [28], Trantina–Johnson model [29], Walker model
3232 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

Fig. 6. da/dN–DK curve.

[30], Forman model [31], and generalized Forman model [32], are commonly used. However, the simplest crack
growth model that describes the crack growth rate under complex spectrum loading is a power law relation [28]:
da n
¼ CðDKÞ ð6Þ
dN
where C and n are constants. The Paris model is still one of the most used expressions for the crack growth
rate, due to its simplicity.
Normally, three regions of crack growth rate are identified as shown in Fig. 6. Region 1 is usually referred
to as the near threshold region due to the threshold stress intensity range, DKth, below which fatigue crack
growth will not occur. This is believed to be true for many materials but for some material-environment com-
binations the slope of the da/dN versus DK relationship has been found to be finite even for growth rates as
low as 107 mm/cycle. Therefore, the stress cycle pertaining to lower stress intensity range than threshold DKth
(shown in Fig. 6) can be regarded as a carrier cycle and should be deleted from the original load history
according to above-mentioned deleting procedure of carrier cycle. Region 2 is usually referred to as the stable
or linear crack growth rate region, since the Paris relation usually fits the data in this region very well. From an
engineering viewpoint, the stable or linear crack growth rate region generally includes a range between
106 mm/cycle and 104 mm/cycle (shown in Fig. 6). Thus, the stress cycle pertinent to the stable or linear
crack growth rate region is deemed as a secondary cycle and the adjacent and sequential secondary cycles
should be merged into a new secondary cycle to shorten the test time by using above-mentioned merging pro-
cedure of secondary cycle. Finally, region 3 is often referred to as the unstable crack growth region, since crack
growth rate increase very rapidly as the maximum stress intensity factor approaches the fracture toughness KC
(shown in Fig. 6). From a fatigue life point of view, the stress cycle pertinent to the unstable crack growth
region is defined as the main cycle and needs to be maintained. By using the quantifying criterion mentioned
above, it is possible to identify all load cycles extracted from an original load history into three kind of stress
cycle: main, secondary and carrier cycles.

3. Quantification criteria to identify load cycle

As crack growth rates were developed for a wider range of rates, it was recognized that the linear relation
(in a log–log scale) represented by Eq. (6) could not describe the crack growth rate for all possible stress
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3233

intensity ranges. The Walker equation, a slight modification to the Paris formulation and one that accounts
for stress ratio effects, for fatigue crack growth is presented [30]
m m2
da=dN ¼ CðDKÞ 1 ð1  RÞ ð7Þ
According to elasto-plastic fracture mechanics [33], the stress intensity factor K can be written as
K ¼ X ðrÞ  Y ðaÞ ð8Þ
where X(r) is a function with respect to stress r, and Y(a) is a function with respect to crack length a. In
general
pffiffiffiffiffiffi
Y ðaÞ ¼ pa  bðaÞ ð9Þ
in which b(a) is the correction coefficient of stress intensity factor for a plate with finite width. By allowing for
the correction for a plastic zone at crack tip, it is possible to obtain
r
X ðrÞ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð10Þ
2
1  paðr=rs Þ
For a plane stress state, a = 1/(2p) and for a plane strain state, a = (1  2m)2/(2p).
Because a fatigue stress cycle is defined by two components namely amplitude Sa and mean Sm, the stress
intensity factor range DK then becomes
DK ¼ K max  K min ¼ ½X ðS max Þ  X ðS min Þ  Y ðaÞ ð11Þ

Since Smax = Sm + Sa and Smin = Sm  Sa, from Eqs. (10) and (11), one has
8 9
>
< >
=
Sm þ Sa Sm  Sa
DK ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Y ðaÞ ð12Þ
>
: 1  pa½ðS m þ S a Þ=rs 2 2>
1  pa½ðS m  S a Þ=rs  ;

Substituting Eq. (12) into Eq. (7) yields


8 9m1
>
< >
=  m2
da Sm þ Sa Sm  Sa 2S a
¼ C qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ½Y ðaÞm1 ð13Þ
dN >
: 1  pa½ðS m þ S a Þ=rs 2 > Sm þ Sa
1  pa½ðS m  S a Þ=rs 2 ;

Letting (da/dN)f = 106 mm/cycle and (da/dN)T = 104 mm/cycle, and substituting (da/dN)f and (da/dN)T
into Eq. (13), it is possible to obtain (a) a filter limit equation to determine a carrier cycle and (b) a threshold
equation to identify a main cycle. Then, the criteria for load cycle identification for accelerated test can be
established as
8 9m1
>
< >
=  S m2
Sm þ Sa Sm  Sa a ðda=dN Þf
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 6 m2 m ð14Þ
>
: r2  paðS m þ S a Þ2 2  paðS  S Þ2 > ; Sa þ Sm 2 C ½rs  Y ða0 Þ 1
s r s m a
8 9m1
>
< >
=  S m2
Sm þ Sa Sm  Sa a ðda=dN ÞT
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi P m2 m1 ð15Þ
>
: r2  paðS m þ S a Þ > S þ S 2 C½r s  Y ða0 Þ
r2s  paðS m  S a Þ ;
2 2 a m
s

where a0 is the initial crack size. From an engineering viewpoint, a0 can be generally chosen as a visible and
detectable macro-crack size of about 1.00 mm, a value chosen on the basis of practical considerations.
If the load cycle extracted by Eqs. (1)–(4) satisfies the inequality (14), then this load cycle can be identified
as a carrier cycle. If the load cycle extracted by Eqs. (1)–(4) satisfies the inequality (15), then this load cycle is
as a main cycle. If a load cycle does not satisfy either of inequalities (14) and (15), then it can be regarded as a
secondary cycle.
3234 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

For a stress cycle of a secondary cycle in the elastic range, it is not necessary to correct for the plastic zone
at crack tip. Thus, Eqs. (12) and (13) become respectively

DK ¼ 2S a Y ðaÞ ð16Þ
da
¼ 2ðm1 þm2 Þ C  S aðm1 þm2 Þ ðS a þ S m Þm2 ½Y ðaÞm1 ð17Þ
dN

a 1.5

3
30

40 90
170

b 220

200

180

160
Nominal Stress (MPa)

140

120

100

80

60

40

20

0
0 5 10 15 20
Time (s)

c 220

200

180

160
Nominal Stress (MPa)

140

120

100

80

60

40

20

0
0 1 2 3 4 5
Time (s)

Fig. 7. (a) The specimen (size unit: mm). (b) Original load history. (c) Generation load history for accelerated tests.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3235

and the inequalities (14) and (15) become


ðda=dN Þf
S ðam1 þm2 Þ ðS a þ S m Þm2 6 ðm1 þm2 Þ m1 ð18Þ
2 C ½Y ða0 Þ
m2 ðda=dN ÞT
S ðam1 þm2 Þ ðS a þ S m Þ P ð19Þ
2ðm1 þm2 Þ C ½Y ða0 Þm1

4. Equivalent damage formulations

Separating the variables of Eq. (17) and integrating, the crack growth life under the stress cycle (Sa, Sm) of
secondary cycle in the elastic range can be expressed as
m Z acr
ðS a þ S m Þ 2 m
N ¼ ðm þm Þ ðm þm Þ ½Y ðaÞ 1 da ð20Þ
2 1 2 CS a 1 2 a0
Substituting Eq. (9) into Eq. (20) gives
m2
ðS a þ S m Þ
N ¼Q ð21Þ
S ðm
a
1 þm2 Þ

with
R acr m1
m1
a0
a 2 ½bðaÞ da
Q¼ ðm1 þm2 Þ m1 ð22Þ
2 p C
2

where acr is the critical crack length. The latter can be determined using the following equation as
Kc
acr  bðacr Þ ¼ pffiffiffi ð23Þ
ðS m þ S a Þ p
It is seen that Eq. (21) describes the Sa  Sm  N fracture surface in three-dimensional coordinate system. The
Sa  Sm  N surface is also termed as the generalized S  N surface.
According to the Palmgren–Miner rule [1,2], the damage resulting from a stress cycle of (Sa, Sm) is
1
DðS a ; S m Þ ¼ ð24Þ
N ðS a ; S m Þ
where N(Sa, Sm) is determined from Eq. (21). From Eqs. (21) and (24), the equivalent damage of a new merged
stress cycle comprising two adjacent and sequential secondary cycles of ðS a1 ; S m1 Þ and ðS a2 ; S m2 Þ shown in Fig. 4
can be determined as

Table 1
Experimental lives of LY12 aluminum alloy specimens
Specimen number Original load history Generated load history
(cycles) (blocks) (cycles) (blocks)
1 249,603 1653 58,940 1592
2 227,708 1508 52,133 1409
3 256,851 1701 57,830 1562
4 226,500 1500 71,891 1943
5 206,568 1368 59,681 1613
6 56,425 1525
7 56,017 1513
8 59,239 1501
Mean life 1546 1594.75
Total test time 21 h 37 min 8 h 45 min
Mean test time 4 h 19 min 1 h 5 min
3236 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

ðm1 þm2 Þ ðm þm Þ ðm þm Þ
ðS a Þeq S a1 1 2 S a2 1 2
m ¼ m þ m ð25Þ
½ðS a Þeq þ ðS m Þeq  2 ðS a1 þ S m1 Þ 2 ðS a2 þ S m2 Þ 2

500
b
400
Nominal Stress (MPa)

300

200

100

-100

0 20 40 60 80
Time (s)

c 500

400
Nominal Stress (MPa)

300

200

100

-100

0 5 10 15 20
Time (s)

Fig. 8. (a) The specimen (size unit: mm). (b) Original load history. (c) Generation load history for accelerated tests.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3237
ðm1 þm2 Þ ðm1 þm2 Þ
S S
Letting (Smin)eq = Smin2 = Sm2  Sa2, H ¼ ðS a1a1þS m1 Þm2 þ ðS a2a2þS m2 Þm2 , from Eq. (25), it is possible to obtain
m 1
ðS a Þeq ¼ ½H  ðS a2 þ S m2 Þ 2 m1 þm2 ð26Þ
and
m 1
ðS m Þeq ¼ S m2  S a2 þ ½H  ðS a2 þ S m2 Þ 2 m1 þm2 ð27Þ

5. Experimental verification

In order to verify the criteria for load cycle identification and the generation principles for accelerated load
histories, three kinds of specimens made of LY12 aluminum alloy, 40CrNiMoA and 30CrMnSiNi2A alloyed
steels were used in fatigue comparative tests between the original and accelerated load histories.
Test 1 concerns LY12 aluminum alloy specimens of a shape and size shown in Fig. 7a. The test is carried
out on an MTS-880-50KN fatigue testing machine at loading frequency of 15 Hz under room temperature and
atmospheric conditions to verify the generation approach of load history for accelerated test. The original load
history is shown in Fig. 7b. From the literature [34], the da/dN–DK curve of LY12 aluminum alloy is obtained
as
da
¼ 1:19  102 ðDK Þ3:83 ð1  RÞ1:43 ð28Þ
dN
pffiffiffiffi
with DK th ¼ 2:8 MPa m at the stress ratio of R = 0.05.
By means of Eqs. (1)–(4), (18), (19) and (28), all load cycles can be extracted from the original load history
given in Fig. 7b and identified into the carrier, secondary and main cycles. According to the procedure shown
in Fig. 5, the carrier cycles are deleted, the secondary cycles are merged based on Eqs. (26)–(28), and then a
new load history for accelerated test is generated as shown in Fig. 7c. From Fig. 7b and c, it can be ascertained
that there are 151 load cycles in a block of the original load history, and 37 cycles in a block of new generation
load histories. In Fig. 7b and c, all carrier cycles are deleted, the secondary cycles are merged largely, and the
main cycle and the sequence between main and secondary cycles are maintained.
Five and eight specimens are used for fatigue testing using the original and accelerated load histories,
respectively, and experimental results are shown in Table 1. From Table 1, it is found that the total test time
for five specimens under the original load history is 1297 min (or 21 h and 37 min) and the mean test time for
every specimen is 259 min (or 4 h and 19 min). The total test time for eight specimens under the accelerated
load history is 525 min (namely 8 h and 45 min) and the mean test time for every specimen is 65.6 min (or 1 h
and 5 min). The relative deviation in the mean life of the accelerated test from the original load history test is
about j1594:751546j
1594:75
 100% ¼ 3:15%, and the average saved test time for every specimen is 193.4 min (3 h and

Table 2
Experimental lives of 40CrNiMoA alloyed steel specimens
Specimen number Original load history Generated load history
(cycles) (blocks) (cycles) (blocks)
1 1,091,937 1793 213,173 1533
2 834,330 1370 203,770 1465
3 984,758 1617 168,739 1213
4 841,635 1382 216,563 1558
5 878,765 1442 169,031 1216
6 760,587 1247 206,972 1489
7 194,739 1401
8 220,988 1589
9 177,781 1279
Mean life 1475.17 1415.89
Total test time 149 h 46.7 min 49 h 13 min
Mean test time 24 h 57.8 min 5 h 28 min
3238 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

13 min). This implies adequate close agreement between the shortened or accelerated test programme and the
original or extended test programme for engineering application.
Test 2 concerns 40CrNiMoA alloyed steel specimens of a shape and size shown in Fig. 8. All specimens
have an initial prefabricated crack of 0.5 mm through linear cutting and polishing. The tests are carried
out on an MTS-880-500KN fatigue testing machine at a loading frequency of 10 Hz under room temperature
and atmospheric pressure conditions. As in the previous test, the original load history is shown in Fig. 8b.

Fig. 9. (a) The specimen (size unit: mm). (b) Original load history. (c) Generation load history for accelerated tests.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3239

From the literature [34], the da/dN–DK curve of 40CrNiMoA alloyed steel at the stress ratio of R = 0.1 is
obtained as
da 2:95
¼ 1:56  104 ðDKÞ ð29Þ
dN
pffiffiffiffi
with DK th ¼ 5:54 MPa m.
Again, by means of Eqs. (1)–(4), (18), (19) and (29), all load cycles can be extracted from the original load
history given in Fig. 8b and identified into the carrier, secondary and main cycles. A similar process is followed
for the load history for the accelerated test programme shown in Fig. 5 resulting in the extraction of the car-
rier, secondary and main cycles in Fig. 8c. From Fig. 8b and c, it can be deduced that there are 609 load cycles
in a block of the original load history, and 139 cycles in a block of the accelerated test load history.
Six and nine specimens are used for fatigue testing using the original and accelerated load histories, respec-
tively, and experimental results are shown in Table 2. From Table 2, it is found that total test time for six
specimens under the original load history is 8986.69 min (or 149 h and 46.7 min) and the mean test time
for every specimen is 1497.78 min (or 24 h and 57.8 min). The total test time for nine specimens under the
accelerated load history is 2952.9 min (or 49 h and 13 min) and the mean test time for every specimen is
328.1 min (or 5 h and 28 min). The relative deviation in the mean life of the accelerated crack propagation test
from the original load history test is about j1475:171415:89
1475:17
j
 100% ¼ 4:02%, and the average saved test time for
every specimen is 1169.68 min (19 h and 30 min). Again, it is evident that the new approach using accelerated
test data gives results that agree very well with the actual test programme for engineering application.
Test 3 concerns 30CrMnSiNi2A alloyed steel specimens of a shape and size shown in Fig. 9. The tests are
carried out on an MTS-880-500KN fatigue testing machine at a loading frequency of 10 Hz under room tem-
perature and atmospheric pressure conditions. As in the previous test, the original load history is shown in
Fig. 9b. From the literature [34], the da/dN–DK curve of 30CrMnSiNi2A alloyed steel is obtained as
da
¼ 3:011  108 ðDKÞ2:45 ð1  RÞ0:98 ð30Þ
dN
pffiffiffiffi
with DK th ¼ 3:67 MPa m at the stress ratio of R = 0.1.
By means of Eqs. (1)–(4), (18), (19) and (30), all load cycles can be extracted from the original load history
given in Fig. 9b and identified into the carrier, secondary and main cycles. A similar process is followed for the
load history for the accelerated test programme shown in Fig. 5 resulting in the extraction of the carrier, sec-
ondary and main cycles in Fig. 9c. From Fig. 9b and c, it can be shown that there are 6906 load cycles in a
block of the original load history, and 1663 cycles in a block of the accelerated test load history.
Three and five specimens are used for fatigue testing using the original and accelerated load histories,
respectively, and experimental results are shown in Table 3. From Table 3, it is calculated that the total test
time for the three specimens under the actual load spectra is 1820 min (or 30 h and 20 min) and the mean test
time for every specimen is 607 min (or 10 h and 6 min). The total test time for the five specimens under the
accelerated load spectra is 728 min (or 12 h and 8 min) and the mean test time for every specimen is
121 min (or 2 h and 1 min). From Table 3, it is evident that the relative deviations in the mean value of fatigue
life under the accelerated spectra compared to that under the actual spectrum test is j53:2552:31 53:25
j


Table 3
Experimental lives of 30CrMnSiNi2A alloyed steel specimens
Specimen number Original load history Generated load history
(cycles) (blocks) (cycles) (blocks)
1 367,340 53 87,307 50
2 297,193 43 65,194 37
3 427,556 61 115,307 66
4 106,088 60
5 62,739 35
Mean life 52.3 53.25
Mean test time 10 h 6 min 2 h 1 min
3240 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

Fig. 10. (a) Full-scale fatigue test of helicopter tail. (b) Fatigue load application points for full-scale test of helicopter tail. (c) Fatigue load
spectrum in transverse direction. (d) Fatigue load spectrum in vertical direction. (e) Fatigue crack occurrence site. (f) Fatigue crack growth
curve.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3241

f T h

1200

1000

800

600

400

200

-200
20 30 40 50 60 70 80 a mm

Fig. 10 (continued)

100% ¼ 1:82%. The saved test time for every specimen is about 451 min (7 h and 31 min). Again, it is evident
that the new approach using accelerated test data gives results that agree very well with the actual test pro-
gramme for engineering application.

6. Application in full-scale fatigue test of helicopter tail

A full-scale fatigue test was carried out on a helicopter tail (shown in Fig. 10a) at loading frequency of
15 Hz under room temperature and atmospheric conditions. Fatigue load spectra in transverse and vertical
directions were applied at tail rotor as shown in Fig. 10b to model the actual flight loads of tail rotor. Using
the finite element method (FEM), the critical sections and hazardous locations can be searched and found
from the actual measured load spectra. Subsequently, by means of Eqs. (1)–(4), (18) and (19) and fracture per-
formances of material of hazardous location, all load cycles can be extracted from the actual measured load
history and identified into the carrier, secondary and main cycles. According to the procedure shown in Fig. 5,
the carrier cycles were deleted, the secondary cycles were merged based on Eqs. (26) and (27) and fracture per-
formances of material of hazardous location, and then the new load histories for accelerated test in transverse
and vertical directions were generated as shown in Fig. 10c and d. During fatigue test, all critical sections and
hazardous locations responses were monitored and observed at various test times to find fatigue crack
3242 J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243

occurrence. After 10,512 blocks of load history (or 22,506 flight hours), an S-shape crack with a length of
32 mm appears at the skin surface of bottom helicopter tail near the right region of tail-lamp shade (shown
in Fig. 10e). Fatigue crack growth curve with experimental time were observed and recorded as shown in
Fig. 10f. The full-scale fatigue test has spend an experimental time of 3658.9 h (or 11,253 blocks of load history
or 22,506 flight hours) until the crack reaches a length of 84.5 mm.

7. Conclusions

The focus of this paper has been to present new generation approach of load history for full-scale acceler-
ated tests. Explicit criteria for load cycle identification and equivalent damage calculation are presented. The
applicability of the new approach has been shown for three validation examples. Good agreement of exper-
imental lives is achieved between the original and generated load histories using the newly developed
approaches. Finally, the generation approach of load history is applied to full-scale accelerated fatigue test
of helicopter tail, demonstrating the practical and effective use of the proposed approach.

Acknowledgement

This project was supported by the National Natural Science Foundation and the Aeronautics Science
Foundation in China, and the Engineering and Physical Sciences Research Council in UK.

References

[1] Miner MA. Cumulative damage in fatigue. J Appl Mech 1945;12:159–64.


[2] Palmgren A. Die Lebensdauer von Kugellagern. Z Vereins Deut Ingenieure 1924;68:339–41.
[3] Matsuishi M, Endo T. Fatigue of metals subjected to varying stress. In: Proceedings of the Kyushu branch of Japan society of
mechanics engineering, Fukuoka, Japan, 1968. p. 37–40 [in Japanese].
[4] Stephens RI, Dindingert PM, Gunger JE. Fatigue damage deleting for accelerated durability testing using strain range and SWT
parameter criteria. Int J Fatigue 1997;19:599–606.
[5] Aicher W, Branger J, Van Dijk GM, Ertelt J, Hück M, De Jonge JB. Description of a fighter aircraft loading standard for fatigue
evaluation FALSTAFF. Common Report of F + W Emmen, LBF, NLR, IABG, 1976.
[6] Mitchenko EI, Prakash RV, Sunder R. Fatigue crack growth under an equivalent FALSTAFF history. Fatigue Fract Engng Mater
Struct 1995;18(5):583–95.
[7] Schütz D, Lowak H, De Jonge JB, Schijve J. A standardised load sequence for flight simulation tests on transport aircraft wing
structures. LBF-report, B-106, NLR-Report TR 73, 1973.
[8] Sieg J, Schijve J, Padmadinata UH. Fractographic observations and predictions on fatigue crack growth in an aluminium alloy under
MiniTWIST flight-simulation loading. Int J Fatigue 1991;13(2):139–47.
[9] Schütz W, Klätschke H, Hück M, Sonsino CM. Standardized load sequence for offshore structures-WASH I. Fatigue Fract Engng
Mater Struct 1990;13(1):15–29.
[10] Fowler KR, Watanabe RT. Development of jet transport airframe fatigue test spectra. In: Potter JM, Watanabe RT, editors.
Development of fatigue loading spectra. ASTM STP 1006. Philadelphia: American Society for Testing and Materials; 1989. p. 36–64.
[11] Wanhill RJH, Hart WGJ, Schra L. Flight simulation and constant amplitude fatigue crack growth in aluminum–lithium sheet and
plate. In: Proceedings of the 16th international committee on aeronautical fatigue symposium, Tokyo, Japan, May 1991.
[12] Schütz W. Standardized stress–time histories—an overview. In: Potter JM, Watanabe RT, editors. Development of fatigue loading
spectra. ASTM-STP 1006. Philadelphia (PA): American Society for Testing and Materials; 1989. p. 3–16.
[13] Heuler P, Klätschke H. Generation and use of standardised load spectra and load–time histories. Int J Fatigue 2005;27(8):974–90.
[14] Schijve J, Vlutters AM, Ichsan A, Kluit JCP. Crack growth in aluminium alloy sheet material under flight-simulation loading. Int J
Fatigue 1985;7(3):127–36.
[15] Yan JH, Zheng XL, Zhao K. Experimental investigation on the small-load-omitting criterion. Int J Fatigue 2001;23(5):403–15.
[16] Schön J. Spectrum fatigue loading of composite bolted joints—small cycle elimination. Int J Fatigue 2006;28(1):73–8.
[17] Dowling NE. Estimation and correlation of fatigue lives for random loading. Int J Fatigue 1988;10:179–85.
[18] Socie DF, Artwohl PJ. Effect of history editing on fatigue crack initiation and propagation in a notched member. In: Bryan DF,
Potter IM, editors. Effect of load history variables on fatigue crack initiation and propagation. ASTM STP 714. Philadelphia: Amer-
ican Society for Testing and Materials; 1980. p. 3–23.
[19] Pompetzki TH, Topper TH, DuQuesnay DL. The effect of compressive underloads and tensile overloads on fatigue damage
accumulation in SAE 1045 steel. Int J Fatigue 1990;12:207–13.
[20] Klemenc J, Fajdiga M. An improvement to the methods for estimating the statistical dependencies of the parameters of random load
states. Int J Fatigue 2004;26(2):141–54.
J.J. Xiong, R.A. Shenoi / Engineering Fracture Mechanics 75 (2008) 3226–3243 3243

[21] Xiong JJ, Shenoi RA. An integrated and practical reliability-based data treatment system for actual load history. Fatigue Fract
Engng Mater Struct 2005;28(10):875–89.
[22] Anthes RJ. Modified rainflow counting keeping the load sequence. Int J Fatigue 1997;19:529–35.
[23] Rychlik I. Simulation of load sequences from rainflow matrices: Markov method. Int J Fatigue 1996;18(7):429–38.
[24] Amzallag C, Gerey JP, Robert JL, Bahuaud J. Standardization of the rainflow counting method for fatigue analysis. Int J Fatigue
1994;16:287–93.
[25] Klemenc J, Fajdiga M. A neural network approach to the simulation of load histories by considering the influence of a sequence of
rainflow load cycles. Int J Fatigue 2002;24(11):1109–25.
[26] Miller MS, Gallagher GP. An analysis of several fatigue crack growth rate descriptions. Measurement and data analysis, ASTM STP
738, 1981. p. 205–51.
[27] Hoeppner DW, Krupp WE. Prediction of component life by application of fatigue crack growth knowledge. Engng Fract Mech
1974;6:47–70.
[28] Paris PC, Erdogan F. A critical analysis of crack propagation laws. J Basic Engng, Trans ASME (Ser D) 1963;85:528–34.
[29] Trantina GG, Johnson CA. Probabilistic defect size analysis using fatigue and cyclic crack growth rate data. Probabilistic fracture
mechanics and fatigue methods, ASTM STP 798, 1983. p. 67–78 .
[30] Walker EK. The effect of stress ratio during crack propagation and fatigue for 2024-T3 and 7075-T6 aluminum. Effects of
environment and complex load history on fatigue life, ASTM STP 462, 1970. p. 1–14.
[31] Forman RG et al. Numerical analysis of crack propagation in cyclic loaded structures. J Basic Engng, Trans ASME (Ser D)
1967;89:459–65.
[32] Newman Jr JC. A crack opening equation for fatigue crack growth. Int J Fract 1984;24:131–5.
[33] Hwang KC, Yu SW. Elastic–plastic fracture mechanics. Beijing: Hsinghua University; 1985.
[34] Gao ZT, Jiang XT, Xiong JJ, et al. Experimental design and data treatment for fatigue performance. Beijing: Beihang University
Press; 1999.

S-ar putea să vă placă și