different fatigue dynamics under statistically and spectrally similar

8
Different Fatigue Dynamics Under Statistically and Spectrally Similar Deterministic and Stochastic Excitations Son Hai Nguyen, Mike Falco, Ming Liu, and David Chelidze * Department of Mechanical, Industrial and Systems Engineering University of Rhode Island, Kingston, RI 02881 Estimating and tracking crack growth dynamics is essential for fatigue failure prediction. A new experimental system— coupling structural and crack growth dynamics—was used to show fatigue damage accumulation is different under chaotic (i.e., deterministic) and stochastic (i.e., random) loading, even when both excitations possess the same spectral and statistical signatures. Furthermore, the conven- tional rain-flow counting method considerably overestimates damage in case of chaotic forcing. Important nonlinear loading characteristics, which can explain the observed discrepancies, are identified and suggested to be included as loading parameters in new macroscopic fatigue models. Keywords: material fatigue, chaotic forcing, stochastic forc- ing, damage accumulation, fatigue dynamics, nonlinear char- acteristics 1 Introduction The fatigue life prediction in variable amplitude loading conditions has received considerable attention, since many structures in service are not subject to constant amplitude cyclic loads. It is currently understood that load factors are nonlinearly coupled with the fatigue life [1, 2] (for example, altering the application sequence of the same large and small amplitude loads results in different fatigue lives). Thus, the most widely used method, Palmgren-Miner rule [3–5], which is based on a superposition principle, is inadequate in prac- tice. Even, when the applied load is stationary over time, the coupling between the structural dynamics and fatigue evolu- tion can drastically alter their dynamics [6]. Thus, it is im- portant to identify essential nonlinear load factors that con- tribute to fatigue life. In the next section, a novel fatigue testing rig based on inertial forces is described. The new testing rig has three * Address all correspondence to this author. Email: [email protected] Accepted for publication in Applied Mechanics on 07/09/2013. major advantages over traditional apparatuses. In particular, it has ability to: (1) mimic complicated real-world load his- tories, such as chaotic and random excitations; (2) conduct fatigue tests at various R-ratios 1 (including zero or negative), which is controlled accurately and can be adjusted during a test; and (3) conduct high frequency tests (going a little higher than 30 Hz based on the current design). In our experiment, fatigue tests are conducted using chaotic (deterministic) and random (stochastic) load time- histories. Both loads have similar power spectral densities and probability distribution functions, but different tempo- ral structures. The mean and variance of the excitations are kept constant throughout the experiments. Thus, superposi- tion based cycle counting methods should be applicable. The rain-flow counting method and Palmgren-Miner rule are ap- plied to the crack load time histories to quantify the fatigue damage variables. Nonlinear loading characteristics are pro- posed to explain the differences in the results, followed by a discussion and conclusions. 2 New Fatigue Testing Apparatus 2.1 Mechanical setup The experimental system was first introduced in [7]. A picture and schematic of the test rig are shown in Fig. 1. The mechanical backbone of the system is a slip table guided by four linear bearings on two parallel rails mounted to a granite base. An LDS V722 electromagnetic shaker is used to drive the slip table under various loading conditions. The spec- imen is a single edge notched beam which is simply sup- ported by pins on each end. Two inertial masses, guided by linear bearings on a central rail mounted to the slip ta- ble provide dynamic loads. The masses are kept in contact with the specimen with the use of two pneumatic cylinders. Different R-ratios can be obtained by adjusting the pressure 1 R = σ min /σ max , where σ min is the minimum peak stress and σ max is the maximum peak stress

Upload: lytram

Post on 03-Jan-2017

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Different Fatigue Dynamics Under Statistically and Spectrally Similar

Different Fatigue Dynamics Under Statisticallyand Spectrally Similar Deterministic and

Stochastic Excitations

Son Hai Nguyen, Mike Falco, Ming Liu, and David Chelidze∗Department of Mechanical, Industrial and Systems Engineering

University of Rhode Island, Kingston, RI 02881

Estimating and tracking crack growth dynamics is essentialfor fatigue failure prediction. A new experimental system—coupling structural and crack growth dynamics—was usedto show fatigue damage accumulation is different underchaotic (i.e., deterministic) and stochastic (i.e., random)loading, even when both excitations possess the samespectral and statistical signatures. Furthermore, the conven-tional rain-flow counting method considerably overestimatesdamage in case of chaotic forcing. Important nonlinearloading characteristics, which can explain the observeddiscrepancies, are identified and suggested to be included asloading parameters in new macroscopic fatigue models.

Keywords: material fatigue, chaotic forcing, stochastic forc-ing, damage accumulation, fatigue dynamics, nonlinear char-acteristics

1 IntroductionThe fatigue life prediction in variable amplitude loading

conditions has received considerable attention, since manystructures in service are not subject to constant amplitudecyclic loads. It is currently understood that load factors arenonlinearly coupled with the fatigue life [1, 2] (for example,altering the application sequence of the same large and smallamplitude loads results in different fatigue lives). Thus, themost widely used method, Palmgren-Miner rule [3–5], whichis based on a superposition principle, is inadequate in prac-tice. Even, when the applied load is stationary over time, thecoupling between the structural dynamics and fatigue evolu-tion can drastically alter their dynamics [6]. Thus, it is im-portant to identify essential nonlinear load factors that con-tribute to fatigue life.

In the next section, a novel fatigue testing rig based oninertial forces is described. The new testing rig has three

∗Address all correspondence to this author. Email: [email protected] for publication in Applied Mechanics on 07/09/2013.

major advantages over traditional apparatuses. In particular,it has ability to: (1) mimic complicated real-world load his-tories, such as chaotic and random excitations; (2) conductfatigue tests at various R-ratios1 (including zero or negative),which is controlled accurately and can be adjusted duringa test; and (3) conduct high frequency tests (going a littlehigher than 30 Hz based on the current design).

In our experiment, fatigue tests are conducted usingchaotic (deterministic) and random (stochastic) load time-histories. Both loads have similar power spectral densitiesand probability distribution functions, but different tempo-ral structures. The mean and variance of the excitations arekept constant throughout the experiments. Thus, superposi-tion based cycle counting methods should be applicable. Therain-flow counting method and Palmgren-Miner rule are ap-plied to the crack load time histories to quantify the fatiguedamage variables. Nonlinear loading characteristics are pro-posed to explain the differences in the results, followed by adiscussion and conclusions.

2 New Fatigue Testing Apparatus2.1 Mechanical setup

The experimental system was first introduced in [7]. Apicture and schematic of the test rig are shown in Fig. 1. Themechanical backbone of the system is a slip table guided byfour linear bearings on two parallel rails mounted to a granitebase. An LDS V722 electromagnetic shaker is used to drivethe slip table under various loading conditions. The spec-imen is a single edge notched beam which is simply sup-ported by pins on each end. Two inertial masses, guidedby linear bearings on a central rail mounted to the slip ta-ble provide dynamic loads. The masses are kept in contactwith the specimen with the use of two pneumatic cylinders.Different R-ratios can be obtained by adjusting the pressure

1R = σmin/σmax, where σmin is the minimum peak stress and σmax is themaximum peak stress

Page 2: Different Fatigue Dynamics Under Statistically and Spectrally Similar

2 3

15 13

5

11 10 9

6 7 81

14 12

4

Fig. 1: Top: experimental apparatus. Bottom: schematic ofthe apparatus. 1. shaker; 2. granite base; 3. slip table; 4.linear bearings for the slip table; 5. back mass; 6. specimensupports; 7. pneumatic cylinder supports; 8. slip table rails;9. front cylinder; 10. front mass; 11. the specimen; 12. linearbearings for the masses; 13. central rail for the masses; 14.back cylinder; and 15. flexible axial coupling.

in each cylinder and the amplitude of the signal supplied tothe shaker. Pressures within the cylinders, however, must belarge enough to keep the masses in contact with the speci-men at all times. This requirement allows for the transfer ofinertial forces to the specimen. When the slip table moves,the specimen drives the masses to follow the motion of theslip table.

The specimens are designed to follow the fracture tough-ness test standard ASTM E1820-08a [8] and their modelis shown in Fig. 2. The specimen is made of 6061aluminum bar stock, with dimensions 314.70 × 20.32 ×6.35 mm (length × height × width), and the fatigue is ini-tiated by a machined ’V’ notch. The notch is cut to a depthof 7.62 mm by a blade 0.762 mm thickness, and with a 60degree V-shaped tip. The initial crack length was examined

Fig. 2: Model of a specimen. The machined notch can beseen in the center, as well as the round hole on the left andthe oblong slot on the right.

under a Stocker and Yale optical micrometer at 40× magni-fication. The crack lengths of all the fatigue tests recordedwere measured to be 7.823±0.069 mm.

In order to get rid of residual and machining stresses,the specimens are fully annealed prior to testing. First, thespecimens are loaded into a heat treating oven, which is pre-heated to 775◦ F and remain there for two hours. Then theoven is shut off and the specimens are allowed to slowly cool.The rate of cooling is kept lower than 50◦ F per hour, whichis achieved by the insulative properties of the oven.

2.2 InstrumentationThe structural response is measured using accelerom-

eters and an eddy current displacement sensor. Two single-axis accelerometers from PCB Piezoelectronics, model num-ber 333B42, are used. The sensitive axis of each accelerom-eter is always kept parallel with the horizontal x-axis. Oneeddy current sensor from Lion Precision, model number U5with an ECL202 driver, is used to measure the relative dis-placement between one of the masses and the slip table.

Drive signals for the testing rig are generated by a pro-gram controlled function generator (Tektronix AFG 30222),which can generate different types of excitation signals basedon either internal functions (such as random and harmonic)or stored arbitrary time histories (such as chaotic). The func-tion generator is connected to a PC using a USB cable in or-der to have full control of the excitation signals. The preloadis measured by two pressure sensors which are directly con-nected to the pneumatic cylinders. The dynamical forces be-tween the pneumatic cylinders and the masses are measuredby two piezoelectric force sensors (Model 208C02 from PCBPiezoelectronics). All data from the sensors are recorded us-ing a data acquisition (DAQ) card from National Instrumenton a PC; and all control and DAQ tasks are implementedthrough a LabView program.

The new design has two major advantages over a simi-lar inertial force based fatigue apparatus described in [9,10].Firstly, the horizontal setup reduces the excitation force nec-essary to drive the system. Therefore, a similar load capac-ity can be achieved by using a smaller, and lower cost elec-tromagnetic shaker. Secondly, the slip table is built usinga standard mounting surface, which allows for simple mod-

Page 3: Different Fatigue Dynamics Under Statistically and Spectrally Similar

ifications to accommodate various types of specimens andconfigurations.

3 Description of Experiment and ResultsIn our experiment, chaotically and stochastically excited

fatigue tests are conducted. Both excitation sources exhibitsimilar spectral and statistical properties, but they are fun-damentally different in their temporal structure. Random orstochastic source is non-deterministic, while chaotic is gen-erated by simulating a deterministic dynamical system. Inparticular, we have used a steady state response of the fol-lowing double-well Duffing oscillator as the chaotic excita-tion signal supplied to the shaker:

x+0.25 x−0.6x+ x3 = 0.2 cos t , (1)

where the over-dot indicates differentiation with respect todimensionless time. The response of the Eq. (1) is sampleduniformly using dimensionless ∆t = 0.2. To generate spec-trally and statistically similar stochastic excitation, the deter-ministic Duffing signal is modified using the method of sur-rogate time series as described in Ref. [11]. In this iterativeprocedure, the chaotic signal and its spectrum’s imaginarypart are randomized, and the power spectrum and the prob-ability distribution function of the resulting stochastic signalis matched with the chaotic signal’s. Fig. 3 shows the com-parison of the resulting histograms and the power spectrumsof the original chaotic and the resulting stochastic signals.

The topologically equivalent phase space of the Duffingattractor is also reconstructed from the recorded x time seriesusing delay coordinate embedding as described in Ref. [12].In particular, a delay time of 8 samples is used to generatethe Duffing phase portrait shown on the right plot in Fig. 4.The same delay was used to reconstruct the correspondingsurrogate time-series phase portrait as shown in the left plotof Fig. 4, which shows the clear difference in time evolutionof these trajectories: the deterministic chaotic signal has anattractor, while stochastic surrogate data does not. These twosignals supplied to the shaker are henceforth called Chaoticand Random. The experimental design is very straightfor-ward. The nominal excitation amplitude is set to keep thetest time manageably short for quick turnaround time. Thisresults in a total of ten sets of data for analysis shown in Table1. It should be noted that all names are indicative of the orderthe test are run and no further inference should be made.

Some of the loads’ metrics are included in Table 1 toshow that the basic statistical qualities of both signals matchwell and do not show any significant deviation from exper-iment to experiment. Any drift in these statistics, from testto test, is due to the small differences between the specimensand/or the lack of a sophisticated shaker output controller.However, these tests have proven very repeatable and the

Fig. 3: Histogram and power spectrum of the original chaotic(left) and stochastic (right) signals

−0.4 −0.2 0 0.2 0.4 0.6

−0.4

−0.2

0

0.2

0.4

0.6

xn

xn+8

−0.4 −0.2 0 0.2 0.4 0.6

−0.4

−0.2

0

0.2

0.4

0.6

xn

xn+8

Fig. 4: Delay coordinate reconstruction of phase portraits forthe stochastic (left) and original chaotic (right) signals

variations in the statistics of the excitation signals are neg-ligible at this amplitude level.

The actual load histograms and power spectral densitiesfrom the slip table acceleration signals (during experiments1-A and 1-B, respectively) are shown in Fig. 5. The accelera-tion signals were collected over ten minutes. The probabilitydistributions for the voltage signals supplied to the shakermatch almost identically (Fig. 3). However, these signalsthen pass through the shaker and couplings to reach the table.The table acceleration probability distributions are not thenentirely identical as seen in Fig. 5, while the power spec-tral densities remain fairly similar for both stochastic andchaotic loads. Furthermore, the differences in the histogramsare only observed for small amplitudes. The correspond-ing phase portraits are reconstructed and shown in Figure 6,where again we see the characteristics observed for the origi-nal input signals. It would be a very difficult task to build in-put surrogate data which would match the table acceleration

Page 4: Different Fatigue Dynamics Under Statistically and Spectrally Similar

Table 1: Test results. Mean, variance, skewness and kurtosis are calculated from table acceleration data. Damage is estimatedbased on the rain-flow counting method and the Palmgren-Miner rule.

Random TTF Damage Mean Variance Skewness Kurtosis

Signal hrs:min D V V2 – –

1–A 0:53 0.97 -0.0020±0.0014 0.0740±0.0038 0.0012±0.0060 2.5850±0.0053

2–A 1:07 1.06 -0.0021±0.0014 0.0716±0.0038 0.0081±0.0049 2.5827±0.0066

3–A 0:44 1.00 -0.0020±0.0015 0.0742±0.0038 0.0092±0.0076 2.5980±0.0094

4–A 0:52 1.07 -0.0019±0.0015 0.0747±0.0041 0.0073±0.0069 2.5898±0.0096

5–A 0:45 0.86 -0.0038±0.0006 0.0724±0.0036 0.0081±0.0052 2.5941±0.0062

Average 0:52 0.99 -0.0023±0.0015 0.0733±0.0018 0.0067±0.0067 2.5892±0.0094

Chaotic TTF Damage Mean Variance Skewness Kurtosis

Signal hrs:min D V V2 – –

1–B 1:44 1.48 -0.0011±0.0014 0.0717±0.0010 -0.0086±0.0052 2.6878±0.0770

2–B 1:50 2.08 -0.0031±0.0014 0.0717±0.0011 -0.0091±0.0062 2.6636±0.0878

3–B 1:46 1.51 -0.0019±0.0015 0.0725±0.0008 -0.0075±0.0068 2.6259±0.0775

4–B 1:42 1.39 -0.0025±0.0015 0.0721±0.0010 -0.0051±0.0041 2.6927±0.0503

5–B 1:13 1.03 -0.0017±0.0014 0.0717±0.0010 -0.0062±0.0050 2.6996±0.0699

Average 1:39 1.50 -0.0021±0.0016 0.0719±0.0010 -0.0074±0.0057 2.6720±0.0785

properties of the chaotic signal. As long as the main factorssuch as mean, and minimum and maximum acceleration arepreserved, the differences are deemed acceptable.

The first and most obvious piece of data that gives anindication of a fatigue process is the time to failure (TTF).A clear trend emerges in the TTF data shown in Table 1:the random excitation is much more damaging to the speci-men than the chaotic excitation. These test results show thatthe system under chaotic excitation outlasts the system un-der Random excitation by nearly a factor of two in all ac-counts. The question then is if these differences can be ex-plained solely by the differences observed in the excitationhistograms shown in Fig. 5. Since the mean, variance, skew-ness and kurtosis of the excitation do not vary significantlyduring the experiment, superposition principle based cyclecounting methods should still account for the differences ob-served in the histograms.

4 Cumulative Fatigue DamageRain-flow counting method [13, 14] and the Palmgren-

Miner rule are used here to quantify the damage based onnumber of stress cycles experienced. Theoretically, the cri-teria for failure would occur when the damage variable,0≤ D≤ 1, reaches one.

The stress at the crack tip is approximated by Euler-Bernoulli beam theory. The specimen is regarded as a simplysupported beam with applied loads, as shown in Fig. 7. P is

−20 −10 0 10 200

0.5

1

1.5

2x 10

4

Acceleration (m/s2)

Num

ber

of O

ccure

nces

0 0.2 0.4 0.6 0.8 1−20

−10

0

10

20

30

Frequency

Pow

er

Spectr

um

Magnitude (

dB

)

−20 −10 0 10 200

0.5

1

1.5

2x 10

4

Acceleration (m/s2)

Num

ber

of O

ccure

nces

0 0.2 0.4 0.6 0.8 1−20

−10

0

10

20

30

Frequency

Pow

er

Spectr

um

Magnitude (

dB

)

Fig. 5: Histogram and power spectrum of the table accelera-tion data of the random (left) and chaotic (right) signals

the static mean force due to the pressure within the cylinders,a is the relative acceleration between the masses and the ta-ble. In this testing, both cylinders have equal pressure. Therain-flow cycle counting method is applied to the time seriesof the maximum stress at the notch of the beam. The objec-tive of this method is to reduce a variable load history signalinto a sequence of constant amplitude cycles. Fig. 8 shows

Page 5: Different Fatigue Dynamics Under Statistically and Spectrally Similar

−15 −10 −5 0 5 10 15−15

−10

−5

0

5

10

15

xn

xn+6

−15 −10 −5 0 5 10 15−15

−10

−5

0

5

10

15

xn

xn+6

Fig. 6: Reconstructed phase portrait of the table accelerationdata for the random (left) and chaotic (right) signals

the histograms of cycle amplitudes from stress histories oftest 1-A and 1-B. While, in general, the TTF depends on thesequence in which the different amplitude cycles are applied,in our tests the load profiles are not altered over time, and thesame load sequence is applied repeatedly. Thus, any differ-ences observed in TTF are only attributable to the differencein temporal structure of each load signal.

The Palmgren-Miner rule uses an S-N curve for the testspecimen to evaluate and determine the damage variable.This curve indicates how many stress cycles can a materialexperience at each stress amplitude level. Then the damagevariable corresponding to each amplitude of stress in Fig. 8is the ratio of the actual cycles experienced over the maxi-mum possible as given by the S-N curve. Since D is simplya ratio, its exact value depends on the S-N curve used and isinherently nondescript. A simple S-N curve can be expressedby the equation

log10 S = A+B log10 N , (2)

where S is the stress amplitude and N is the number of cy-cles until the failure at that stress level. For the comparisonof chaotic and random loadings, the S-N curve is chosen toyield D≈ 1 for all random loadings. Hence, the coefficientsA and B are determined to be 2.58 and −0.25, respectively.Then the damage variable is calculated for each test. Theresults are shown in Table 1.

Typically, the Palmgren-Miner rule would be used topredict the failure by characterizing a typical stress historyover some time, and approximating the stress history overthe expected lifetime of the fatiguing structure. The failurecriteria would be met once the damage variable D = 1. Inour analysis, the damage criteria is calculated over the ac-tual stress history of the entire load time history. Therefore,since the failure criteria is equivalent for all tests it can beexpected that for all tests D would be approximately equal ifthe linear damage law is applicable. This is true across testsof the same type, chaotic or random. However, in compari-son between the two, the linear damage law does not reflect

Fig. 7: Beam diagram with applied forces

0 10 20 30 400

0.5

1

1.5

2

2.5x 10

4

σ (Mpa)

0 10 20 30 400

0.5

1

1.5

2

2.5x 10

4

σ (Mpa)

Fig. 8: Histograms of cycle amplitudes counted by the rain-flow counting method in random (left) and chaotic (right)tests.

the observed effects of the excitation signals’ time history onTTF. In particular, it significantly overestimates damage forthe chaotic excitation.

The fatigue testing using random and chaotic excitationthat have nearly identical statistical and spectral propertiesshowed that specimens failed twice as fast during chaoticforcing (Table 1). The actual stress loads at the crack havealmost exactly the same spectral characteristics, but his-tograms show that chaotic forcing has more low amplitudeloads, while maintaining approximately the same loadingfor larger amplitudes (Figure 4). Everything being equaland keeping nominal excitation amplitude constant over thewhole experiments, the linear Palmgren-Miner damage law,which is independent of the loading time-history, shouldhave accounted for the differences in histograms. However,while it was adequate to provide consistent estimates of dam-age during random excitation experiments, Palmgren-Minerrule significantly overestimated the damage variable for thechaotic excitation. This result implies that the excitation’stemporal structure is very important in fatigue accumula-

Page 6: Different Fatigue Dynamics Under Statistically and Spectrally Similar

0 5 10 150

20

40

60

80

100

Embedding dimension (size)

FN

Ns(%

)

0 5 10 150

20

40

60

80

100

Embedding dimension (size)

FN

Ns(%

)Fig. 9: Estimation of embedding dimension for the random(left) and chaotic (right) signals

0 100 200 3002.2

2.4

2.6

2.8

3

3.2

3.4

3.6

n (time steps)

ln(δ

n)

0 100 200 3002.2

2.4

2.6

2.8

3

3.2

3.4

3.6

n (time steps)

ln(δ

n)

Fig. 10: Average trajectory divergence rates for all records inrandom loading (left) and chaotic loading (right)

tion, even when main statistical moments of the load at thecrack do not change significantly, and both linear statisticsand spectral properties of both load signals are very similar.Thus, we need to look at nonlinear characteristics of the ex-citation time series.

5 Nonlinear Characterization of LoadingThe two main features of dissipative nonlinear dynami-

cal systems are Lyapunov exponents [15–17] and fractal di-mensions [18–20]. The Lyapunov exponents are the aver-age exponential rates of separation of trajectories that areinfinitesimally close initially. Fractal dimensions relate tothe complexity of the system’s attractor, and can provide alower bound on the active degrees-of-freedom in a dynamicalsystem. These quantities are characteristics of a determinis-tic dynamical system, and are not defined or appropriate forstochastic systems. Here we will not focus on the estimationof the actual Lyapunov exponents or fractal dimensions, butinstead look at nonlinear features extracted by applying theappropriate algorithms to the load time histories. The hopeis that these features will capture the differences that can beused to explain the discrepancies in the observed TTFs.

In phase space of a dynamical system, two trajectorieswith initial separation δ0 diverge after n time steps as

δn = δ0 eλ tn , (3)

where λ is the Lyapunov exponent, and tn = n∆t. The exis-tence of a positive Lyapunov exponent indicates the chaoticbehavior of a system. We use the algorithm used to calculatethe maximal Lyapunov exponent [12, 21] to obtain featurecurves that describe average divergence of nearby trajecto-ries in the phase space over finite time period.

Since crack propagation can make our system struc-turally unstable [6], stress data is divided into data records of100,000 points in each. A delay time of 7 time steps [22] isused for the delay coordinate embedding of the time series.Embedding dimension is determined for the random signalusing the first minimum of false nearest neighbors (FNN)curve [24]. Zero FNN indicates complete unfolding of theattractor for that dimension, which ensures the uniqueness ofthe resulting reconstructed trajectory. Fig. 9 shows that anembedding dimension of 5 is optimal for both signals. How-ever, the FFN curve never reaches zero for the random signaland continues to increase after the initial dip. While the ini-tial dip in the FNN trend suggests the existence of short-timecorrelations in the random data (as in colored noise), laterincrease indicates its infinite dimensionality (i.e., there is nofinite embedding dimension for which the embedded trajec-tories become unique).

The average trajectory divergence is estimated in eachdata record. The results for test 1-A and 1-B are shown inFig. 10. As expected, there is a clear difference betweenthe divergence rates observed for the chaotic and random ex-citations for small time scales, and these features are fairlyconsistent during each experiment. The divergence rate atsmall scale for the random signal is much higher than forthe chaotic, which could also explain differences observed inhistograms for the small amplitudes. For the random excita-tion starting at low amplitudes, large amplitudes are reachedvery fast. In contrast, the chaotic signal’s divergence is moregradual and slower. In other words, time correlation persistslonger in the chaotic time series, which causes more coher-ent structures at small length and time scales oin the chaoticsignal.

Figure 11 shows the average trajectory divergence ratesfor all ten tests in the first data record. These results areconsistent for each test type and again show clear differencein the random and chaotic divergence rates.

The correlation sum is usually used to estimate the frac-tal dimension of an attractor [23]. However, since randomsignal does not have a finite-dimensional attractor, we usethe correlation sum as a feature differentiating our signals.Let a set of points {yk}n

k=1 uniformly sample an attractor.Then the correlation sum C(ε) approximates the cumulativedistribution of distances on the attractor as:

C(ε) =2

(n− s)(n− s−1)

n

∑i=1

n

∑j=i+s+1

Θ(ε−

∥∥yi−y j∥∥)

∼ εD2 ,

(4)

Page 7: Different Fatigue Dynamics Under Statistically and Spectrally Similar

0 100 200 3002.2

2.4

2.6

2.8

3

3.2

3.4

n (time steps)

ln(δ

n)

0 100 200 3002.2

2.4

2.6

2.8

3

3.2

3.4

n (time steps)

ln(δ

n)

Fig. 11: Average trajectory divergence rates for all tests usingrandom loading (left) and chaotic loading (right)

−3.5 −3 −2.5 −2 −1.5 −1 −0.5 0−4

−3

−2

−1

0

1

2

3

4

〈log10(ǫ)〉

log 1

0(C

(ǫ))

Correlation Sum

Random

Chaotic

Fig. 12: Correlation sum for all ten tests

where Θ is a Heavyside step function, n is the total numberof points, s is the time interval needed to remove any timecorrelations in the pair of points, and D2 is the estimated cor-relation dimension. The modified algorithm given in [20] isused to estimate the correlation sum for the different lengthscales ε, and results are shown in Fig. 12. Again we see thatfor small length scales there are clear differences. For thechaotic signal there are considerably more smaller distancesin the data than for the random signal. However, both sig-nals have identical correlation sum for the large amplitudesor distances, which again correlates well with the similarityof the histograms for large amplitudes.

6 DiscussionDifferences in excitations observed in histograms did

not prove sufficient to provide consistent estimates of TTFusing Palmgren-Miner rule and Rain-Flow counting method.In particular, they overestimated damage for the chaotic ex-citation by almost a factor of two. Additional nonlinearcharacteristics (trajectory divergence rates and correlationsum) showed more drastic differences in the small time- andlength-scale dynamics that can explain the observed differ-

ences in TTF. Larger divergence rates for small scales in therandom signal clearly correlate well with shortened TTF inthe random tests. For small length- and time-scales, theselarger rates cause the load on the crack to change faster forthe random signal than for the chaotic. Therefore, the crackunder random load experiences fewer low stress cycles whencompared to the chaotic load which can explain the differ-ences in the TTF. In particular, this suggests that fatigue lifedepends not only on the cycles of maximum stress experi-enced, but also on the rate at which this maximum stressis achieved—faster time-rate-of-change of stress being moredamaging. The observed differences in divergence rates arealso reflected in the deviations observed in the correlationsums at small scales. Therefore, both of these features corre-late well with the observed TTF and may be used as loadingparameters for a new damage model that could account forthe observed differences.

7 Conclusion

In this paper, we reported on the experimental resultsfrom fatigue testing in which chaotic and random loadingtests were conducted using a novel fatigue testing apparatus.Both loadings shared the same statistical and spectral prop-erties. However, experimental results showed that randomloading was more damaging than chaotic. Standard, linearPalmgren-Miner rule and rainflow counting method grosslyoverestimated damage for the chaotically excited system,while they were adequate to describe damage in randomlyexcited system. Nonlinear dynamical characteristics of theexcitation time series were considered to explain the differ-ences. These metrics clearly showed that while the signalshad similar characteristics at large scale, for a small scaletrajectory divergence rates were much higher in the randomsignal. In addition, there was more finer structure/detail inthe chaotic signal as shown by the correlation sum. These ob-servations stipulated that crack was experiencing fewer lowstress cycles for the random excitation, which was also re-flected in the excitation histograms. These results also sug-gested that fatigue life depends on the rate at which maxi-mum stress is reached at the crack tip. Therefore, both of theconsidered nonlinear characteristics correlate well with theobserved differences in the TTF, and may be used as loadingparameters for the new damage laws providing more consis-tent TTF estimates.

8 Acknowledgments

This paper is based upon work supported by the NationalScience Foundation under Grant No. 0758536.

Page 8: Different Fatigue Dynamics Under Statistically and Spectrally Similar

References[1] M. Skorupa. Load interaction effects during fatigue

crack growth under variable amplitude loading - a lit-erature review. part i: Empirical trends. Fatigue Fract.Engrg. Mater. Struct, 21:987–1006, 1999.

[2] M. Skorupa. Load interaction effects during fatiguecrack growth under variable amplitude loading–a litera-ture review. Part II: Qualitative interpretations. FatigueFract. Engrg. Mater. Struct, 22:905–926, 1999.

[3] M. Miner. Cumulative damage in fatigue. Journal ofApplied Mechanics, 67:A159–A164, 1945.

[4] A. Palmgren. Die Lebensdauer von Kugellagern. Ver-fahrenstechinik, Berlin, 68:339–341, 1945.

[5] E. Macha, T. Lagoda, A. Nieslony and D. Kardas. Fa-tigue life under variable-amplitude loading accordingto the cycle-counting and spectral methods. MaterialsScience, 42(3):416–425, 2006.

[6] C.-H. Foong, E. Pavlovskaia, M. Wiercigroch andW. F. Deans. Chaos caused by fatigue crack growth.Chaos, Solitons and Fractals, 16:651–659, 2003.

[7] Michael Falco, Ming Liu, and David Chelidze. A newfatigue testing apparatus model and parameter identifi-cation. ASME 2010 International Design EngineeringTechnical Conferences and Computers and Informa-tion in Engineering Conference, 5(44137):1007–1012,2010.

[8] ASTM E1820-08a. Standard Test Methods for Mea-surement of Fracture Toughness. Annual Book of ASTMStandards. American Society for Testing and Materi-als, Philadelphia, PA., 2008.

[9] Chee-Hoe Foong, Marian Wiercigroch, and WilliamF.Deans. Novel dynamic fatigue-testing device: designand measurements. Measurement Science and Technol-ogy, 17:2218–2226, 2006.

[10] Nikola Jaksic, Chee-Hoe Foong, Marian Wiercigroch,and Miha Boltezar. Parameter identification and mod-elling of the fatigue-testing rig. International Journalof Mechanical Sciences, 50(7):1142–1152, 2008.

[11] Thomas Schreiber and Andreas Schmitz. Improvedsurrogate data for nonlinearity tests. Phys. Rev. Lett,77:635–638, 1996.

[12] H. Kantz and S. Schreiber. Nonlinear Time Series Anal-ysis. Cambridge University Press, 2004.

[13] S.D. Downing and D.F. Socie. Simple rainflow count-ing algorithms. International Journal of Fatigue,4 (1):31 – 40, 1982.

[14] ASTM E 1049-85. Standard practices for cycle count-ing in fatigue analysis. ASTM International.

[15] J. B. Dingwell. Lyapunov exponents. Wiley Encyclo-pedia of Biomedical Engineering, 2006.

[16] A. Wolf, J. Swift, H. Swinney, and J. Vastano. Deter-mining lyapunov exponents from a time series. PhysicaD, 16 (3):285–317, 1985.

[17] A. Wolf. Quantifying chaos with lyapunov exponents.

Chaos, 273–290, 1986.[18] P. Grassberger, and I. Procaccia. Characterization of

strange attractors, Physical review letters, 50 (5):346–349, 1983.

[19] P. Grassberger. Generalized dimensions of strange at-tractors. Physics Letters A, 97 (6):227–230, 1983.

[20] P. Grassberger, and I. Procaccia. Measuring thestrangeness of strange attractors. Physica D. 9 (1):189–208, 1983.

[21] Michael T. Rosenstein, James J. Collins, and CarloJ. De Luca. A practical method for calculating largestLyapunov exponents from small data sets. PHYSICAD, 65:117–134, 1993.

[22] Andrew M. Fraser and Harry L. Swinney. Independentcoordinates for strange attractors from mutual informa-tion. Phys. Rev. A, 33:1134–1140, Feb 1986.

[23] James Theiler. Spurious dimension from correlation al-gorithms applied to limited time-series data. PhysicalReview A, 34:2427–2432, 1986.

[24] Matthew B. Kennel, Reggie Brown, and Henry D. I.Abarbanel. Determining embedding dimension forphase-space reconstruction using a geometrical con-struction. Phys. Rev. A, 45:3403–3411, Mar 1992.