Вы находитесь на странице: 1из 25

FRACTURE, FATIGUE AND MECHANICAL RELIABILITY

PART II

An Introduction to Fatigue

by

James W. Provan

Spring 2008 Department of Mechanical Engineering University of Victoria Victoria, BC, Canada

Part II-1

1.

Introduction to the study of fatigue

Fatigue, the loss of strength or other important property as a result of cyclic stressing over a period of time, is a general phenomenon occurring in most materials. In all walks of life, components fail under fluctuating loads that are well below their static design loads. Such failures have been recognized for centuries with the first known fatigue research, relating to the failure of conveyor chains used in Clausthal coal mines, being published by Albert in 1829 [1]. The term fatigue has origins dating back to 1839, when Poncelet first described metallic structures as being tired or worn out [2, 3]. Another early fatigue researcher was Whler, who in 1860 conducted fatigue investigations of railroad axles and created the concept of the fatigue endurance limit [4]. His tests on railway axles led to the widely accepted stress-life, S-N or Whler curves discussed in Chapter 3. Early construction, such as those made of bricks or wood, was mainly designed to withstand compressive stresses. During the industrial revolution, however, these materials were partially replaced by wrought and ductile steels and cast iron. Understandably this lead to unexpected structural failure at stresses well below their anticipated levels. Fractures of railway axles, tracks, boilers and associated pressure vessels were so common that in the 1870's the British magazine Engineering printed weekly statistics on railway accidents [5]. One of the most famous early fatigue failures occurred in 1919 in Boston, USA. A large tank, 15m high and 27m in diameter, ruptured and over 7.5 million litres of molasses were spilled. Without warning the sides burst apart killing 12 and injuring 40 [3]. During World War II the United States manufactured liberty ships, which had an all-welded hull instead of the traditional riveted one. In 1943, one of the Liberty ships broke in two while sailing between Alaska and Siberia. Soon it was discovered that other Liberty ships suffered from similar fractures. Out of the 2700 ships built, 400 had fractures; in 20 ships the failure was total and 10 of them broke completely in two. Fatigue cracks were able to run right through the entire hull, as no riveted seams prohibiting their progress [3]. In the 1950s, de Havilland Comet jet airliners suffered a series of unexpected mid-air loses [3, 8]. Originally, Comets had square passenger windows at the corners of which cracks formed as a result of variations in cabin pressure. These cracks grew to their then unknown critical size prior to catastrophically propagated through the entire length of the main cabin. Whler's S-N curves may be used to determine the fatigue endurance of components under, primarily, constant amplitude loading. However, in practice, constant amplitude loading is quite rare. Many of the current methods of calculating damage under variable amplitude loading are based on Palmgrens [6] and Miners [7] cumulative linear damage summation theory. The damage caused by a single cycle in a variable amplitude loading is assumed to be linear and equal to the damage caused by a cycle of the same size under constant amplitude loading, and thus may be determined with the help of an S-N curve. Palmgren and Miners rule may be considered the first fatigue life prediction method. Fatigue of metals by the slow growth of microscopic flaws was documented by Ewing and Humfrey, even though the mathematical tools for quantitative treatments of fracture had not yet been developed [8]. The fatigue life of a component was understood to be composed of three

Part II-2

parts, namely crack initiation, crack propagation and final fracture. At that time, much of the research emphasis shifted towards the investigation of crack propagation. The previously introduced pioneering study of Griffith, who developed a quantitative connection between fracture stress and flaw size, was established [9]. Griffiths approach only applies to ideally brittle materials, but was modified in 1948 by Irwin and 1949 by Orowan, and since has been applied in the fatigue investigation of many materials [10, 11]. As discussed in Part I, the revolutionary concept of the stress intensity factor was introduced in 1957 by Irwin [12]. In essence K was developed as a scale factor to define the magnitude of the crack tip stress field. In 1961, Paris, Gomez and Anderson were the first to suggest that the increment of fatigue crack advance per stress cycle could be related to the range of the Irwin's stress intensity factor [13]. At that time their theory did not receive a warm welcome, as faith in the S-N approach was and remains, strong. As is shown in Section 4.4, Paris together with Erdogan in 1963 unambiguously established the validity of their formula, i.e., the growth of a crack under cyclic loading is governed by the Paris-Erdogan law [14]. Forsyth [15] presented a fatigue crack scenario that is still accepted as an indication of how a crack initiates, propagates and then leads to final failure. A schematic of this interpretation is shown in figure 1.1. As discussed by Provan [16], a fatigue failure initiates with a small crack that is minute and even difficult to locate using powerful replication and inspection techniques. For materials whose microstructure is relatively homogeneous, this stage of the fatigue process is usually shear stress dependent and acts on microscopic slip planes that are preferentially oriented to this stress. Characteristically, this crack starts at a point of discontinuity in a material, such as a defect, machining mark or a change in cross-section. Once the crack has initiated, which according to Forsyth corresponds to the end of the Stage I portion of the fatigue process, the stress concentration effect becomes predominant and the crack progresses in a plane perpendicular to the local principle normal stress. The rate at which this Stage II crack propagates is accepted as being crack length dependent thereby lending itself to analyses based on LEFM concepts.

The inability to distinguish between Stage I and Stage II processes in actual design situations contributes to the uncertainty that exists in fatigue analyses and partially motivates the discussion contained in Part III of these notes.
As the stressed area decreases in size, the stress magnitude increases until, finally, the remaining area catastrophically fails. Crack initiation/nucleation is a process that leads to the development of a microcrack. This process is most susceptible to the influence of the local microstructure. The mechanisms of the initiation process have been affiliated with microscopic flaws at grain boundaries, twin boundaries, inhomogeneities (both microstructural and compositional), inclusions and stress concentrations at the microscopic and macroscopic levels [17]. Fatigue failures are most often initiated at a surface of a component due to the numerous stress concentrations found there. These include scratches, dents, machining marks, roughness due to corrosion and surface offsets, where different glide planes have undergone different amounts of net slip under cyclic strain. Within the material the initiation of fatigue cracks is associated with intermetallic inclusions, pores or soft spots in the microstructure [18]

Part II-3

Figure 1.1

The various modes of fatigue failure

There is no simple or clear delineation of the boundary between fatigue crack initiation and propagation. Basically, after microcrack formation the influence of the local microstructure decreases and the crack grows at an increasing rate under repeated external forces. The direction of the crack growth is normal to the applied maximum principal stresses [18]. In the final fracture phase the structure is spontaneously ruptured because the remaining cross section is too small to resist the external forces. Previously it was believed that the initiation phase covers the main part, say 50 to 75%, of the fatigue life of a structure. More recent studies, however, have revealed that microcracks may be generated after the first 1% of the lifetime [19]. Fracture mechanisms in metals and alloys are usually divided into the following categories: brittle and ductile fracture (including cleavage and intergranular fracture), discussed in Part I and fatigue, discussed here. The entire process of fatigue failure is a combination of fatigue and either ductile or brittle final fracture. During the crack propagation period the fatigue mechanism is understandably predominant, whereas the final fatigue fracture may be based upon either brittle or ductile mechanisms, depending on the material, temperature, loading rate and constraints [2]. Visual inspection of a fatigue fracture surface reveals two zones: the smoother fatigue propagation zone and rougher zone of final fracture. In a detailed examination the existence of ripple markings on the smooth zone may be observed. These markings are known as striations. They are the small ridges that lie perpendicular to the direction of crack propagation. When the strain at the tip of the crack reaches a critical value it brings about a ductile tear mechanism, which together with the subcritical crack extension leaves fatigue striations, also known as beach markings, on the fracture surface [18]. Forsyth [17] first proved that each striation corresponds to one load cycle. In general, the stage II striations are better defined in ductile rather than brittle materials. In

Part II-4

aluminum alloys the striations are precisely positioned, continuous lines, whereas in stronger steels they tend to be discontinuous, short and more randomly positioned. Due to the variation of stress triaxiality in thicker plates, (see Part I), a crack tends to grow faster at the center of the plate, causing a tunneling effect. On plate surfaces the crack grows in a ductile plane stress manner, while the interior adopts a more brittle plane strain characteristic. The crack exhibits a tunneling effect and produces so called shear lips. Near the free surface the crack growth occurs at 45 from the maximum principal stress, as shown in figure 1.2.

Figure 1.2

Shear lip formation

The fatigue life of a component is, then, the sum of time spent in these three mentioned phases. Crack initiation life, N i is defined by the number of loading or straining cycles required to develop a microcrack. N p is the number of cycles required to propagate a crack to some critical dimension. The final fracture phase forms a negligible portion of the fatigue life, so the total fatigue life, N T , may be expressed as:

NT = N i + N p

(1.1)

A wide range of mechanical, structural and environmental factors have an effect on the crack initiation and propagation processes. Thus it is understandable, that even minor variations in, for example, the definition of crack initiation period will result in a variety of life predictions. A materials scientist is likely to be concerned with microscopic processes that determine the crack initiation from flaws along persistent slip bands. The mechanical engineer, however, is likely to define crack initiation based on the available crack detection equipment and is usually limited by the fraction of a millimeter resolution of the equipment used [2]. Classical approaches to fatigue problems, introduced in Chapter 3, involve the determination of so called total fatigue life in terms of applied stress range, strain range, mean stress and environment. This method makes no distinction between crack initiation and growth, but embodies the two into a single formulation that predicts the total life to catastrophic failure.

Part II-5

Whlers stress-life approach together with the Palmgren-Miner linear cumulative damage rule, fall into this particular category. In the 1950s the above mentioned aircraft accidents prompted the development of fatigue methods specifically for aircraft design. The safe-life approach, developed from fatigue tests to predict the replacement time for aircraft components, quite often involved very conservative safety (or ignorance) factors. Most components were designed so that their replacement times exceeded the expected service life of the entire aircraft, with some critically loaded components requiring inspection and replacement if necessary. Once a component reaches its replacement time, its safe-life is reached and it is retired, whether or not any fatigue cracks are present. According to the safe-life approach a component is always replaced before it develops a fatigue crack. The fail-safe approach to fatigue design was introduced in 1960s. The goal in this approach is to design multiple load path structures, so that if an individual element fails, the remaining elements should have sufficient structural integrity to enable the structure to operate until the crack is detected. In addition to multiple load paths, structures designed using this approach typically contain crack stoppers to prevent undesirable crack growth. Fail-safe approaches call for periodic inspections to enable prompt repair or replacement strategies [19, 20]. In the early 1970s a new method of fatigue design was introduced, namely the damage tolerance approach discussed in Part III. This approach is based upon fracture mechanics techniques, the basic assumption being that all engineering components contain flaws. During an inspection, existing flaws may be detected in components and the number of cycles to failure determined by using crack growth equations. In order to determine the inspection interval the number of cycles to failure is then usually divided by a factor to ensure that each crack is inspected at least once before the crack theoretically propagates to failure [19]. The damage tolerance approach involves the calculation of crack growth rates by the laws of fracture mechanics, which limits the application possibilities of this approach. According to LEFM, the damage tolerance approach is applicable where predominantly elastic loading conditions prevail and where the crack tip plastic zone is relatively small compared to the component and crack size [21]. Fatigue failure under constant amplitude loading (CAL) generally forms the basis of fundamental studies. While constant amplitude stress-life, strain-life, and crack growth rate properties have contributed much to the understanding of fatigue behavior, materials selection, and life predictions, in practice many engineering components are subjected to complex load histories. As a result, classic fatigue analyses fail to predict the fatigue crack propagation accurately. Since the 1960's researchers have been able to show that the variation in the magnitude of the load cycles may significantly affect the fatigue life of components. The development of reliable life prediction models capable of handling complex variable amplitude load histories is one of the toughest challenges of fatigue research. There have been models based on compressive residual stresses, yield zones, strip-yielding and stress-strain hysteresis loops, to mention a few. These models give similar results, but there is no general consensus on any particular model at this time. What has been established is that two important concepts govern crack growth rates under variable amplitude loading: crack tip plasticity and crack closure. These processes are discussed in Section 4.5.

Part II-6

References:

[1] [2] [3] [4] [5] [6] [7] [8]

[9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21]

Frost, N.E., Marsh K.J, Pook L.P., Metal Fatigue, Clarendon Press, (1974). Suresh, S., Fatigue of Materials, Cambridge University Press, New York, (1991). Hertzberg, R.W., Deformation and Fracture Mechanics of Engineering Materials, John Wiley & Sons, (1976). Whler, A., Engl. Abstr. Eng. vol 2, (1871). Rossmanith, H.P., (ed), Fracture Research in Retrospect: An Anniversary Volume in honor of Professor George R. Irwin's 90th Birthday, A.A. Balkema, (1998). Palmgren, A., Die Labensdauer von Kugellagern, VDI-Z. vol 68, pp 339-341, (1924). Miner, A.M., Cumulative Damage in Fatigue, J. Appl. Mech., pp 151-164, (1945). Smith, R.A., Thirty Years of Fatigue Crack Growth - A Historical Review in Fatigue Crack Growth 30 Years of Progress. Proceedings of a Conference on Fatigue Crack Growth, Cambridge, UK, 20th September, (1984). Griffith, A.A., Trans., Royal Society of London, vol 221, (1920). Reprinted with additional commentary, Trans. ASM, vol 61, p 871, (1968). Irwin, G.R., Fracturing in Metals, ASM, p 147, (1949). Orowan, E., Fatigue and Fracture of Metals, MIT Press, p 139, (1950). Irwin, G. R., Analysis of Stresses and Strains Near the End of a Crack Traversing a Plate, J. Appl. Mech., vol. 24, pp 361-364, (1957). Paris, P. C., Gomez, M. P., and Anderson, W. P., A Rational Analytic Theory of Fatigue, The Trend in Engineering, vol. 13, pp 9-14, (1961). Paris, P. C., and Erdogan, F., A Critical Analysis of Crack Propagation Laws, Transactions of the ASME, series D, J. Basic Engrg, vol. 85D (4), pp 528-534, (1963). Forsyth, P.J.E., Fatigue Damage and Crack Growth in Aluminum Alloys, Acta. Meta., vol 11, p 703, (1963). Provan, J.W., The Micromechanics of Fatigue Crack Initiation, Modeling Problems in Crack Tip Mechanics, J.T. Pindera, (ed), pp 131-154, (1984). Forsyth, P.J.E., A two-stage process of fatigue crack growth, Proceedings, Crack Propagation Symposium, vol. 1., Cranfield College of Aeronautics, pp. 76-94, (1962). Barsom, J.M., Rolfe, S.T., Fracture and Fatigue Control in Structures, Applications of Fracture Mechanics, Second Edition, Prentice Hall, (1987). Kaplan, M.P., Wolff, T.A., Life Extension and Damage Tolerance of Aircraft, Fatigue and Fracture, ASM Handbook vol 19, ASM International, pp 557-565, (1996). Bruhn, E.F., Analysis and design of flight vehicle structures, Jacobs Publishing, (1973). Ikonen, K., Kantola, K., Murtumismekaniikka, p 844, Otatieto, Finland, (1986).

Part II-7

2.

Fatigue cycle counting and simulation procedures


(I am indebted to Mrs. Maarit White for help in writing this Chapter.)

The first mandatory step in any fatigue design analysis is to establish the expected load history. When dealing with variable amplitude loads, cycle counting therefore plays an extremely important role. All such procedures produce a summary from recorded data so that the signal may subsequently be simulated during the subsequent fatigue design process. In the case of fatigue, the recorded data is usually load, stress or strain versus time, while the simulation bears as close a resemblance as possible to the original spectrum. Such spectra are essential both in the experimental fatigue testing of materials and components and in the development of fatigue life prediction methods. The initial fatigue studies of crack growth under service simulated load histories were in the aircraft industrial sector. A typical loading spectrum acting on an aircraft structural component is shown in figure 2.1 [22]. Each flight may be divided into parts, such as taxiing, take-off, climb, cruise, decent, landing and final taxiing. Each part has its own level of periodic or deterministic loads coupled with transient and random (stochastic) loads induced by ever changing circumstances, such as runway surface roughness and air turbulence, etc. The transient loads are usually sudden and short lived, while the stochastic loads are experienced as smaller random oscillations superimposed on the deterministic levels. The important point is that the recorded data and their simulations be firmly established as being representative and repeatable.

Figure 2.1

Example of loading history on lower skin of an aircraft during 1 flight

Any cycle counting algorithm should be able to extract from the recorded data all the information relevant to the fatigue behavior of the component. The three most important

Part II-8

features are the variations in stress amplitude and mean stress, and the sequence of their fluctuations. Currently and as noted in subsequent sections of these notes, the reproduced spectra are being employed successfully in a variety of S-N and theoretical crack growth models to predict fatigue lives. They are also being used in experimental fatigue testing where the aim is to reproduce fatigue damage bearing a close resemblance to that experienced by in-flight loading. Since these tests are usually accelerated tests, where the loads must closely mirror the service loads, the accuracy of both the cycle counting and simulation technique must leave no doubt as to the validity of the predictions.

2.2

Cycle counting algorithms:

At least 10 separate techniques have been and are being used to analyze random data. These methods may be classified into four groups, namely, the level crossing, range/mean, rainflow and probability density function (PDF) methods.
2.2.1 Level crossing method

The level crossing method was successfully introduced in the 1960s by Gassner [23]. Its popularity is due mainly to the ease of data collection and graphical interpretation. In this method, the range between the minimum and maximum signal is divided into arbitrary amplitude levels set at equal increment levels above and below the mean. As shown in figure 2.2, when the signal trace crosses one of these levels a count is registered, with the restriction that only increasing crossings are registered at levels on or above the mean and only decreasing crossings at levels below the mean. The counts at each level result in a statistical summary of stress excursions rather than a record of individual cycles or a sequence of changes in stress amplitude. As a result, the method does not give detailed information on the actual loading spectra being experience by the component in question.
2.2.2 10% reset method

By eliminating the small reversals in a signal trace, the 10% reset method was able to overcome some of the weaknesses in the level crossing method. It uses the same signal but with a different counting procedure, i.e., a trace logs the level it crosses but does not register it until the trace passes through a predetermined reset level, in this case 10% of the maximum signal. This effectively eliminates the trace excursions that are not associated with crossing the reset level. In addition to the problems associated with the above level crossing method, the 10% reset method may also eliminate some large excursions that have the potential of greatly influencing fatigue damage estimates.

Part II-9

Figure 2.2 2.2.3

Level crossing cycle counting method

Range/mean method

The range/mean method analyzes fluctuations in both stress amplitude and mean stress [24]. The signal is divided into half cycles (valley-peak or peak-valley) and their stress amplitude and mean are recorded. Unfortunately, this method exhibits weaknesses similar to the previous two. The low amplitude, high frequency noise that is superimposed on a high amplitude, low frequency signal results in a series of small half cycles with a gradually changing mean. Since this has the potential of underestimating the severity of the high amplitude signal, while the almost negligible damage due the small cycles is included, this method has not gained acceptance.
2.2.4 Rainflow counting

In 1968 Matsuiski and Endo [25] introduced a cycle counting method, which separates high and low amplitude cycles and records them in a physically meaningful way by reducing complex loading data into a series of threshold nominal stresses. It has universally become the accepted method of both counting cycles and simulating reconstructed spectra. There exist two different descriptions of rainflow counting, which look different but are known to give the same result. The first method documented in figure 2.3 is visually clear and explains why this method was given its name. The illustrated short stress-time history with its time axis pointing down may be treated as if it was a section through a pagoda with rain flowing down every roof starting at the highest point and dripping off each extremity. A flow stops if it:

Part II-10

a. b. c.

reaches the end of the signal, joins water dripping down from a higher extremity, or comes opposite to a larger peak of the same type as that at which it started.

Horizontally, the length of each terminated flow is recorded as a half-cycle stress range. Referring to figure 2.3, the first half-cycle starts at 1 and stops opposite 5 (criterion c), giving it an amplitude of 1 - 4. The half-cycle starting at 2 stops opposite 4, resulting in a 2 3 amplitude. The half-cycle starting at 3 terminates when joining the 'water' dripping down from reversal 2 (criterion b), giving it a 3 2 amplitude and forming a pair with the previous halfcycle. Continuing the process, results in three single half-cycles and three equal pairs.

Figure 2.3

Rainflow counting, pagoda description

(There exists a similar counting method known as the range-pair method. This method counts a range as a cycle if it can be paired with a subsequent range of equal magnitude in the opposite direction. Occasionally, especially in the case of a complicated history, some ranges without their counterparts are counted as cycles, as are some ranges that are interrupted by smaller ranges.)

Part II-11

The second and recommended description of the rainflow method is called the four-point [24, 26] algorithm, which scans the signal and extracts a cycle as soon as three consecutive ranges verify the following relationship. 1. Four successive points 1, 2, 3 and 4 are chosen from a signal history and their stresses, 1, 2, 3 and 4 recorded. 2. Three consecutive stress ranges are formed from these recorded stresses: 1 = |2 - 1|, 2 = |3 - 2| and 3 = |4 - 3 |. 3. If 2 1 and 2 3, then the cycle represented by 2 and 3 (its extreme values) is extracted, the two points 2 and 3 are discarded and the loose ends connected. 4. If 2 1 or 2, then the first point is discarded and a new test performed on points 2, 3, 4 and 5. This test is repeated until all the points in the sequence have been tested. Once the last point has been reached the resulting form is known as a residue. The results as discussed below are stored in a matrix giving the mean value, the stress or strain range and the number of cycles. Viewed graphically, four-point rainflow counting is applied to the stress history shown in figure 2.4i. Figure 2.4ii shows the extraction of the first three cycles, the two middle points in each case being removed and the counting continuing with the reconnected signal. The extraction of two more cycles is illustrated in figure 2.4iii, with the residue of half-cycles, which no longer meet the above criteria, being shown in figure 2.4iv. This residue may either be stored for later sequence reconstruction or be processed further by replicating the residue after itself as shown in figure 2.5 and applying the four-point method to the combined sequence. After processing the final residue may be decomposed for fatigue life predictions or used in sequence reconstruction. While counting is progressing, the extracted cycles are recorded in a rainflow matrix [27]. For example, the stress history extractions of figure 2.4ii are placed in the appropriate cells of the rainflow matrix illustrated in figure 2.6. In this way, the extracted cycle as a pair of nominal stress values (i , j ) is placed in a rainflow matrix in the position (i , j). This process is carried out for the entire stress history. The resulting rainflow matrix will have a diagonal line of zeros and does not contain any information on the sequence of loads. It together with the residue provide sufficient information for reproduction of the original spectrum.

Part II-12

Figure 2.5

Replication of two residues

Part II-13

Figure 2.6 2.2.5

Construction of rainflow matrix

Modified rainflow counting methods

While the standard rainflow counting method has been widely used for many years, it does have shortcomings that were pointed out by Hong [28]. For example, in the low cycle fatigue situation discussed in Chapter 5, Hong suggests a modification that correctly accounts for the closed hysteresis loops that occur in this situation. Furthermore, in attempting to address the sequencing deficiency of the rainflow counting method, Newman [29] introduced a crack-closure (see Section 4.4) based method where the damage is considered to be caused only by the maximum load. Research in finding a way of correctly accounting for sequencing effects is ongoing. Until this is established, the more common applications of the rainflow counting algorithms make full use of the Palmgren-Miner [6, 7] cumulative linear damage rule discussed in the Chapter 3. Finally, another important matrix built into fatigue cycle simulation algoritms, is the transition matrix, which is different from the rainflow matrix [27]. In a transition matrix, all the level transitions of the stress history are placed in the (i, j) position according to the nominal stress values (i , j ) that they represent. Considering figure 2.7, which illustrates a similar load profile to that depicted in figure 2.4 with the exception that the last stress returns to the level of the first (in this case 3), a proper transition matrix must satisfy the property that the sum of cycle counts in each cell of a row above the diagonal must equal the sum of cycle counts in each cell of the column opposite and below the diagonal. This is illustrated in figure 2.8
2.2.6 Probability density function methods

As may be observed from the previous sections, the process of counting cycles is time consuming and requires a reasonably long history for results to be reliable. Several researchers, however, have proposed methods based upon probability density functions instead of rainflow algorithms. For example, marine classification society rules governing the fatigue design of

Part II-14

vessels [30, 31] utilize only two spectral properties to describe the appropriate Weibull PDFs. Other examples will be appropriately introduced throughout the remainder of Part II and Part III.

Figure 2.7

Construction of a transition matrix

Figure 2.8 In a transition matrix the sum of a row and column on the opposite sides of the diagonal are equal.

2.3

Simulation algorithms:

Fatigue loading simulation algorithms are used to generate realistic in-service loading spectra that may be used in the experimental testing of specimen, components or whole structures and in estimating theoretical fatigue lives, most often under VAL conditions. For the most part, simulations usually use the results of the cycle counting to reconstruct the desired spectra by employing random number generators. Although the originally developed simulations divided
Part II-15

the loading spectra into blocks and processed these blocks to attain estimates of the corresponding fatigue lives, practically all modern simulation algorithms make use of Markov processes, matrices or chains developed on the basis of rainflow analyses.
2.3.1 Markov processes

A process is called Markovian if the probability of an outcome only depends on the event immediately preceding it, and not on the entire history. Since each point in a fatigue data spectrum is either a minimum or maximum, whose precise nature may be determined from the position of the immediately preceding reversal, the selection of extrema values from such spectra is a Markov process. Inherently, all the data required for an application of Markov processes to reconstruct a loading spectrum are contained in the preceding section's transition matrix. Instead of using the actual accumulated number in each cell of a transition matrix a more useful way to examine the data is by using the transition probability matrix, which is formed by dividing the accumulated number in each cell by the appropriate sum of either all the accumulated numbers in the row above the diagonal or all the numbers in the row below the diagonal. An example is shown in Table 2.1.

Table 2.1

Conversion of transition matrix to transition probability matrix.

As examples and according to this transition probability matrix, a load level transition starting from level 1, a minimum, has a 66.67% chance of moving to level 3. However, for a load level starting from 3, the previous transition must be known so that the direction of this transition may be determined. If the previous transition was from below and the current position is a maximum, then there is a 97.56% chance of moving to level 2 and a 2.44% chance that it will be level 1. Alternatively, if the previous transition was from above and the current position is a minimum, then the chances of moving to levels 4, 5 and 6 are 69.25, 30.47 and 0.28%, respectively. These 3 probabilities illustrate a classical discrete Markov process, or Markov chain. Hence, the transition probability matrix or Markov matrix contains all the necessary information for the reconstruction of a simulation spectrum based upon the generation of random numbers.

Part II-16

2.3.2

Random number generators

Random number generators are deterministic algorithms that produce numbers according to specified distributions. All random number sequences may be generated by randomly selecting elements from a number array. For example, RANMAR [32] generates a 32-bit sequence of floating point random numbers uniformly (Gaussian) distributed on the interval (0, 1), with the end points excluded. With reference to figure 2.8, these values are multiplied be a factor, the factor in this example being 6. By choosing the number of generated points to be 50, the load transitions between levels 1 and 6 result in a fatigue load spectrum similar to that shown in figure 2.9. The corresponding transition and probability transition matrices are given in Table 2.2. When the number of generated random numbers is set to 1000, the two matrices shown in Table 2.3 illustrate smoother characteristics.

Figure 2.9

A random number generated load spectrum [32]

Table 2.2

Transition and probability matrices associated with N = 50

Part II-17

Table 2.3 2.3.3

Transition and probability matrices associated with N = 1000

Monte Carlo (MC) procedures

Monte Carlo (MC) procedures are often used to evaluate random variables governed by complicated probability density functions. Most are associated with random number generators that generate uniform (Gaussian), statistically independent values on the interval [0, 1). MC algorithms are available for a wide selection of probability density functions, the most common being the normal, gamma, binomial and Poisson distributions. Incorporating a MC technique known as the acceptance-rejection or von Neumann method, it has been established that good simulated fatigue load spectra based upon Markovian transition matrices may be generated.
2.3.4 Exceedance diagrams

In some cases it is beneficial to express the results of cycle counting in the form of exceedance diagrams, which show the number of records for which the load values exceed specified values [33], see figure 2.10. The construction of an exceedance diagram starts by determining the exceedance steps of the original in-service fatigue data and counting the corresponding exceedance values, marked by crosses in figure 2.10. A curve is then drawn through these data points and approximated by a suitable number of load levels; in this illustration, 15 levels both above and below 0. All of the above basic simulation techniques may be modified to suit the specific needs of a particular industrial sector. Some of these are briefly discussed in the next Section.

Part II-18

Figure 2.10

An exceedance diagram

2.4

Standard spectra:

In a growing number of industrial sectors the development and use of standard spectra has become mandatory. Table 2.4 lists the more common [34].
Year 1973 1976 1977 1979 1983 1987 1987 1990 1990 1990 1990 1991 Table 2.4 Name Description of load Transport aircraft lower wing skin Fighter aircraft lower wing skin Random loading Shortened version of TWIST Helicopter main rotor blades Tactical aircraft composite wing skin Cold fighter aircraft turbine disks Hot fighter aircraft turbine disks Offshore, wave action standard history Car components Steel mill drives Horizontal axis wind turbine blades

TWIST FALSTAFF GAUSSIAN miniTWIST HELIX/FELIX ENSTAFF Cold TURBISTAN Hot TURBISTAN WASH CARLOS WALZ WISPER/WISOERX

Standard load spectra [34]

Part II-19

As mentioned previously, the use of these spectra has enabled comparative tests being implemented in the areas of material selection, surface treatment, structural design, etc., and in the acceptance of advanced fatigue life prediction models. In practice these spectra are large files containing load sequences in numerical formats, most often in the form of peak-valley pairs. TWIST, miniTWIST and FALSTAFF are briefly discussed in the following subsections.
2.4.1 TWIST

The Transport WIng STandard (TWIST) load program was developed cooperatively between the Dutch National Aerospace Laboratory (NLR) and the German Laboratorium Fr Betriebstestigkeit (LBF) [35]. It represents the realistic load-time history of the critical lower wing skin at the wing root of a transport aircraft. As illustrated in Table 2.5, for the development of TWIST, the wing root bending moment spectra were obtained from several aircraft types representing a wide range of weights, wing loads, flight distances and cruising speeds. It is worth reviewing the development of TWIST since it illustrates the very considerable international effort required for the development of standard spectra for specific industrial segments. The load spectra for the BAC 1-11, DC-9 and Boeing 737 were calculated from the fight profile shown in figure 2.11 obtained from Lufthansa's short distance operation. Gust load calculations were performed, notably on the Fokker F-27 and DC-10, from data obtained from the British Engineering Sciences Data Unit (ESDU). Taxi and touch down loads for the most part were calculated from information provided by Buxbaum and Gassner, with the Fokker F-27 and DC-10 spectra being calculated by the manufacturer.

Table 2.5

Basic data for the aircraft considered by TWIST [35]

The load spectra of TWIST pertaining to accelerations were developed for civil aircraft in normal airline operations and for Transall military transport from measurements made on a Boeing 720 and Transall research aircraft. The Fokker F-28 spectrum was prepared by the manufacturer and was based upon a restricted level counting method.

Part II-20

Figure 2.11

Lufthansa's short distance mission profile [35]

All aircraft specific spectra were reduced to a basis of 40,000 flights, which is considered an average design life for transport aircraft and normalized to a non-dimensional form by dividing by the mean stress associated with "one-g" cruise conditions. Figure 2.12 shows both the standard TWIST spectrum and those measured from the listed aircraft. The excellent comparison fully justifies the international effort and costs involved with the development of such standard spectra.

Figure 2.12

TWIST load spectra of 40,000 flights for different aircraft [35]

The conversion of this spectrum into experimental test spectra was made by converting the above continuous form into stepped equivalents and defining the sizes of each flight block. This both simplifies the fatigue machine's load controlling system and speeds up the computational treatment

Part II-21

of the spectrum. With reference to figure 2.13, TWIST uses 10 different load levels ranging from a very smooth flight in calm weather, (J), to an extremely rough flight in a thunderstorm, (A), along with 10 cycle blocks. TWIST then periodically repeats each 4,000 flight block.

Figure 2.13

Experimental flight type load spectra [35]

The TWIST fatigue load spectrum file is a 79733 x 10 matrix of load levels, each line consisting of 10 load numbers forming 5 peak-valley pairs. The values have been normalized and the load levels are between 10 and 10. This permits the user to specify the scaling of the data with a linear factor. A common way to scale is to set the maximum load level at a certain fraction of the yield/ultimate strength of the material in question.
2.4.2 miniTWIST

The main concern with TWIST was its absolute size. One flight consists of 100 load cycles and to run 40,000 flights at a frequency of 10 Hz took about one week. Hence in 1979, NLR and LBF introduced miniTWIST, a shortened version of TWIST [36]. The main aim was to minimize the difference between the two spectra. While keeping the number and sequence of the different flight types and magnitude of the load levels the same, the number of cycles per flight was reduced from 100 to about 15 by discarding 95 % of the cycles with the smallest range.
2.4.3 FALSTAFF

The international development effort of NLR, LBF Industrieanlagen-Betreibsgesellshaft (IABG) and the Swiss Federal Aircraft Authority, resulted in the development of the Fighter Aircraft Loading STandard For Fatigue (FALSTAFF) spectrum [37]. TWIST, being focused on gust type loading, was unsuitable for fighter aircraft, which are primarily governed by maneuver

Part II-22

loads. Hence, FALSTAFF focuses on fatigue in the wing root area, governed by overall wing bending and pronounced ground-to-air cycles. FALSTAFF used experimentally measured load factor histories from 342 individual flights of 5 different aircraft as its primary input. The data sources were all given equal representation (40 per aircraft) and the individual load factors all scaled in order to avoid the domination of any one particular history. After the truncation levels were set at the once-per-block exceedance frequency, the load factor spectrum had the appearance shown in figure 2.14.

Figure 2.14

FALSTAFF unified load factor spectrum [37]

The established stress sequence mixture is represented by a bi-variate stress range distribution with the number of discrete stress levels being set at 32. The 200 (5 x 40) different flights were divided into 3 mission groups: flights exhibiting a repetitive pattern of exercises involving severe maneuvering, flights exhibiting severe maneuvering without repetitive patterns of exercises, and the remainder involving moderate or incidental maneuvering. Flight lengths were set to 3 different lengths within each mission group, while the ground based loadings were divided into 2 groups: take-off and landing. Finally the sequence of loads was determined by using a random function generator. The complete FALSTAFF fatigue spectrum consists of 35966 numbers representing 200 flights. Peak and valley values alternate and the spectrum starts with a peak. A widely available version has normalized numbers having a minimum value of 2.4830 and a maximum of 9.3096. This spectrum may be converted to an absolute load range by a multiplicative scale and/or bias factor. The complete transition matrix of the original FALSTAFF spectra is shown in Table 2.6.

Part II-23

Table 2.6 Transition matrix for FALSTAFF [37]

Part II-24

References:

[22]

[23] [24] [25] [26] [27] [28] [29] [30] [31]

[32] [33] [34] [35]

[36] [37]

Skourpa, M., Load Interaction Effect during Fatigue Crack Growth under Variable Amplitude Loading A Literature Review Part I: Empirical Trends, Fatigue and Fracture of Engineering Materials and Structures, vol 21, pp 987-1006, (1998). Gasser, E and Schltz, W., Evaluating vital vehicle components by program fatigue tests, Proceedings 9th Int. Auto. Tech. Congress, I. Mech. Engrg., pp 195-205, (1962). Power, E.M., Cycle counting methods and the development of block load fatigue programs, SAE 780102, (1978). Matsuiski, M. and Endo, T., Fatigue of metals subjected to varying stress, Japan Soc. Mech. Engrg., (1969). Amzallag, C., Standardization of the rainflow counting method for fatigue analysis, Int. J. Fatigue, vol 16, pp 287-293, (1994). Olagnon, M., Practical computation of statistical properties of rainflow counts, Int. J. Fatigue, vol 16, pp 306-314, (1994). Hong, N., A modified rainflow counting method, Int. J. Fatigue, vol 13, pp 465-469, (1991). Newman, J.C., A crack-closure model for predicting fatigue crack growth under aircraft spectrum loading, ASTM, STP 748, pp 53-84, (1981). Fatigue Assessment of Ship Structures, Classification Notes No. 30.7, Det Norske Veritas, Hvik, Norway, September (1998). Fricke, W., Petershagen, H. and Paetzold, H., Fatigue Strength of Ship Structures, Parts I and II, GL-Technology, nos. 1/97 & 1/98, Germanischer Lloyd Aktiengesellschaft, Hamburg, Germany, (1997/98). CERNLIB Short Write-ups, Application Software and Databases: Computing Networks Division, p 327, CERN, Geneva, Switzerland, (1996). Broek, D. and Smith, S.H., The prediction of fatigue crack growth under flight-by-flight loading, Engrg. Fracture Mech., vol 11, pp 123-141, (1979). Schijve, J., Fatigue crack growth under variable amplitude loading, "Fatigue and Fracture", ASM Handbook vol.19, pp 110-132, (1996). de Jonge, J.B., et al., A standardized load sequence for flight simulation tests on transport aircraft wing structures, National Aerospace Laboratory, NLR, TR73029, Amsterdam, Holland, (1973). Lowak, H., et al., miniTWIST, a shortened version of TWIST, Laboratorium Fr Detriebsfestigkeit, TB-146, Darmstad, Germany, (1979). van Dijk, G.M., and de Jonge, J.B., Introduction to a fighter aircraft loading standard for fatigue evaluation FALSTAFF, National Aerospace Laboratory, NLR, MP75017U, Amsterdam, Holland, (1975).

Part II-25

Вам также может понравиться