Вы находитесь на странице: 1из 5

Page 1 of 5

Fatigue User’s Guide > Fatigue Theory > Introduction


Introduction
Fatigue theory for basic stress--life (S-N) and strain-life ( -N) analyses are presented in this chapter. This corresponds to the technical background for Total Life and Crack Initiation (Ch. 5). Theory and
technical information for other analysis types are given in their respective chapters. However a good understanding of the other analysis types is dependent on an understanding of the basic theories
presented here. For this reason, an entire chapter is dedicated to this theoretical discussion.

Background
Purely static loading is rarely observed in modern engineering components or structures. By far, the majority of structures involve parts subjected to fluctuating or cyclic loads. For this reason, design
analysts must address themselves to the implications of repeated loads, fluctuating loads, and rapidly applied loads. Such loading induces fluctuating or cyclic stresses that often result in failure of the
structure by fatigue. Indeed, it is often said that from 80% to 95% of all structural failures occur through a fatigue mechanism.
It is worth noting at the outset that the term fatigue, coined more than a hundred years ago by the French engineer Monsieur Poncelet, may not be the best choice of terminology today, since many
aspects of the phenomenon are distinctly different from the biological counterpart. For example, it is next to impossible to detect any progressive changes in material behavior during the fatigue
process, and therefore failures often occur without warning. Also, periods of rest with the fatigue stress removed do not lead to any measurable healing or recovery of the material. Thus, the damage
done during the fatigue process is cumulative, and generally unrecoverable. From this standpoint, the German term Betriebsfestigkeit (operational strength) is a better descriptor of the phenomenon.
However, since Betriebsfestigkeit involves 17 characters, and fatigue only 7, we shall continue to use the term fatigue!
Fatigue, although a complex subject, has not been neglected by the research community. Estimates indicate that if one wished to keep up to date with all the literature published about fatigue by
reading a paper each working day, one would fall behind by more than a year for each year of reading. Furthermore, attempting to catch up with the backlog would be virtually impossible. Yet the
design analyst and engineer is increasingly challenged by the demands of higher performance, lower weight, and longer life, and all this at a reasonable cost and in as short a time as possible! These
apparently conflicting demands can only be overcome through a consideration of the problems associated with fatigue resistant designs. Up until recently, these problems were summarized as the
following:
• Life calculations are usually less accurate than strength calculations. Order of magnitude errors in life estimates are not unusual.
• Fatigue properties cannot be accurately deduced from other mechanical properties; they need to be measured directly.
• Full-scale prototype testing is usually necessary to assure an acceptable life.
• Laboratory results of tests carried out under identical conditions may differ widely, requiring statistical interpretation.
• Materials and design geometries must often be selected to provide slow crack growth, and if possible, detection of cracks before they become dangerous.
• “Fail-safe” design concepts must often be implemented in order to achieve acceptable reliability. That is, even if a structural element fails, the structure must remain intact and able to support
the loads in the short term.
Modern advances in fatigue life estimation techniques have, to some extent, mitigated these problems. For example, these days it is usual to consider life estimates to be within a factor of two or three
rather than ten. Furthermore, computerized analysis of thousands of laboratory data sets do point to acceptable empirical correlations between monotonic tensile data and fatigue parameters.

The History of Fatigue


For centuries, it has been known that a piece wood or metal can be made to break by repeatedly bending it back and forth with a large amplitude. However, it came as something of a surprise when it
was discovered that repeated loading produced fracture even when the stress amplitude was apparently well below the elastic limit of the material. The first fatigue investigations seemed to have been
reported by a German mining engineer, W. A. S. Albert, who in 1829, performed some repeated loading tests on iron chain. Some of the earliest fatigue failures in service, occurred in the axles of
stage coaches. When railway systems began to develop rapidly in the middle of the nineteenth century, fatigue failures of railway axles became a widespread problem that began to draw attention to
cyclic loading effects. This was the first time that many similar components had been subjected to millions of cycles at stress levels well below the monotonic tensile yield stress. As is often the case
with unexplained service failures, attempts were made to reproduce the failures in the laboratory. Between 1852 and 1870, the German railway engineer August Wöhler setup and conducted the first
systematic fatigue investigation. From this point of view, he may be regarded as the grandfather of modern fatigue thinking. He conducted tests on full-scale railway axles and also on small scale
bending, torsion, and axial cyclic loading specimens for different materials. Some of Wöhler’s data for Krupp axle steel were plotted in terms of nominal stress amplitude versus cycles to failure. This
presentation of fatigue life has become very well known as the S-N diagram. Each curve on such a diagram is still referred to as a Wöhler line.
At about the same time, other engineers began to concern themselves with the problems associated with fluctuating loads in bridges, marine equipment, and power generation machines. By 1900, over
80 papers had been published on the subject of fatigue failures. During the first part of the twentieth century, more effort was placed on understanding the mechanisms of the fatigue process rather than
just observing its results. This activity finally led, in the late 1950s and early 1960s, to the development of two approaches to fatigue life estimation. One method, known as the Manson-Coffin local
strain approach, attempts to describe and predict crack initiation whilst another is based on linear elastic fracture mechanics, LEFM, and was developed to explain crack growth. Most recently, Miller
and his colleagues at Sheffield University, England, have been working on ways of finding a unified theory of metal fatigue, describing crack growth on a microscopic, macroscopic, and structural
level.
From this vast wealth of knowledge, one thing has become clear; modern design analysts and engineers will not create more fatigue resistant components and structures by indulging in more
experimentation, although the need for more research is ever present. From a practical point of view, a more profitable approach is the implementation and efficient use of the knowledge which is
available today.

High Cycle versus Low Cycle Fatigue


Over the years, fatigue failure investigations have led to the observation that the fatigue process actually embraces two domains of cyclic stressing or straining that are distinctly different in character.
In each of these domains, failure occurs by apparently different physical mechanisms: one where significant plastic straining occurs and the other where stresses and strains are largely confined to the
elastic region. The first domain involves some large cycles, relatively short lives and is usually referred to as low-cycle fatigue. The other domain is associated with low loads and long lives and is
commonly referred to as high-cycle fatigue. Low-cycle fatigue is typically associated with fatigue lives between about 10 to 100,000 cycles and high-cycle fatigue with lives greater than 100,000
cycles.
The rules for distinguishing between these two domains are discussed later. For now, it is enough to recognize that remedies for extending fatigue lives in each domain are different. In the high-cycle
fatigue domain, measures such as shot peening and other surface hardening treatments or the use of higher strength materials are beneficial. For low-cycle fatigue, where ductility and resistance to
plastic flow are important, these measures are inappropriate.

Summary
In summary, fatigue analysis may be thought of as a process of initiating and then growing a crack which finally causes the structure to break into two or more pieces. This process can be represented
by the following equation:
Total Life = Crack Initiation + Crack Growth
or

(15-1)

Nf Total fatigue cycles to failure


Ni Number of cycles to initiate an engineering crack
Np Number of cycles to propagate a crack to final fracture
The mathematical models used to simulate the initiation and propagation processes are quite different. The initiation phase is usually modeled using strain-life and cyclic stress-strain curves while the
propagation phase uses crack growth rate versus stress intensity curves.
The original work on fatigue life estimation did not make a distinction between crack initiation and growth. Stress was used as the control parameter and stress-life (or S-N) curves characterized the
material or component response to cyclic loading. This approach was used widely until the 1970s when the new methods started to be used by industry. Unfortunately, there are quite a few inherent
problems with the stress-life approach which cause a large amount of scatter in the experimental fatigue life results. In some companies, due to the extensive use of the stress-life method in past years,
a large quantity of S-N curves and experience have been accumulated. In this case, the stress-life method offers compatibility with previous work and may therefore be more appropriate.
The stress-life approach is still used today for situations which are not easily modeled using the local strain methods, such as welded structures and cast irons. Due to the inherent defects in these
materials which act more like cracks, fatigue damage should really be modeled with LEFM crack propagation tools. Another group of materials which appear to be best modeled using the stress-life
approach are anisotropic and inhomogeneous materials such as composites. However, there are complications even with composites, since the mean stress correction models used in connection with
metals do not apply to all composites.

http://www.mscsoftware.com/training_videos/patran/reverb3/Fatigue%20Users%20Guide/... 5/26/2012
Page 2 of 5

Inputs to Fatigue Life Estimation Models


The fatigue life estimation process requires three main inputs and the process is usually illustrated with a diagram known as the “Five Box Trick” as shown in Figure 15-1.

Figure 15-1 Schematic Illustration of Fatigue Life Estimation


From this diagram, it is easy to see the necessary inputs:
• Materials properties
• Loading in the form of load time histories
• Local stress-strain information from a linear elastic analysis

These inputs are then processed using various fatigue life estimation tools which are described in more detail below. It is important to understand the nature of these inputs so that the fatigue life
estimation is meaningful.

Material Properties
When considering fatigue, it is not sufficient to characterize a material purely in terms of the Young’s modulus and Poisson’s ratio. The chemical composition, heat treatment, and microstructure will
all change the way in which the material responds to cyclic loading. A summary of the materials data types for fatigue life estimation is given below.

Stress-Life Data
For this model, the fatigue response for a material used in a particular construction is the stress amplitude versus log cycles to failure curve. In addition, the Young’s modulus may be required to
convert stress to strain or vice versa. The S-N curve parameters are listed in Table 15-1 and typical S-N curve is shown in Figure 15-2.

Table 15-1 Stress-Life Data Parameter Description

Parameter Name S.J. Units Imperial Units Parameter


Stress range intercept MPa or MN/m2 PSI SRI1
First fatigue strength exponent none none b1
Fatigue transition point cycles none NC1
Second fatigue strength none none b2
exponent MPa or MN/m2 PSI FL
Stress range fatigue limit none none SE
Standard error of Log(N)

Figure 15-2 Example of a Stress-Life Plot

Crack Initiation Data


There are two basic sets of materials properties used for this fatigue life estimation model. They are the cyclic stress-strain curve and the strain-life curve. These are characterized by the parameters in
Table 15-2 with a typical strain-life plot shown in Figure 15-3.

Table 15-2

Parameter Name S.J. Units Imperial Units Parameter


Fatigue strength coefficient MPa or MN/m2 PSI σf'
Fatigue strength exponent none none b
Fatigue ductility coefficient none none εf'
Fatigue ductility exponent none none c
Young’s modulus MPa or MN/m2 PSI E
Cyclic strain hardening exponent none none n'
Cyclic strength coefficient MPa or MN/m2 PSI K'
Cut off in reversals none none Rc

http://www.mscsoftware.com/training_videos/patran/reverb3/Fatigue%20Users%20Guide/... 5/26/2012
Page 3 of 5

Figure 15-3 Example of a Strain-Life Plot

Crack Growth Data


The primary piece of information used in this model is the Paris curve which defines crack growth rate as a function of the stress intensity range. In addition, the fracture toughness and the threshold
stress intensity range is required. These parameters are listed in Table 15-3 with a typical da/dN-Delta K plot in Figure 15-4.

Table 15-3

Parameter Name S.J. Units Imperial Units Parameter


Paris law coefficient m/cycle in/cycle C
Paris law exponent none none m
Delta K threshold at R=0 MPa m1/2 KSI in1/2 D0
Delta K threshold at R-1 D1
Fracture toughness MPa m1/2 KSI in1/2 K1C
Stress ratio at threshold knee MPa m1/2 KSI in1/2 Rc
Stress corrosion threshold none none K1SCC
Unnotched fatigue strength FL
MPa m1/2 KSI in1/2
MPa PSI

Figure 15-4 Example of a Delta K Apparent Plot

Loading
Loading time histories are measurements of loading for a period of time which must be long enough to ensure that the measurement reflects a typical duty cycle. The loading measurements will be
made in the same attitude as the loading applied in the FEA. For example, an axle may experience bending loads in at least two planes together with a torsion load, or a pressure vessel may experience
both pressure and temperature variations in time. Simultaneous measurement of these loads for a typical or worst case event would be made.
Measurement of any physical phenomenon requires a transducer, the electronics to drive the transducer and a recording device. A typical combination would be a load cell, amplifier and FM tape
recorder. Most engineering companies will have some capabilities and experience in doing this kind of work even if the group carrying out the finite element analysis does not. It is important that these
two groups work together in order to achieve the best advantage of the MSC.Fatigue package.
If it is not feasible to measure load or a related parameter, it may be necessary to synthesize a load time history. There are tools in the MSC.Fatigue package to allow you to create, manage, and
graphically edit time histories (PTIME). In some cases) the load variation with time may be trivial and easily created artificially, (e.g., the variation of pressure in a vessel). However, it is important to
remember that many fatigue problems occur because an unexpected load combination occurred. For this reason, it is a truism that “a measurement is worth a thousand guesstimates.”

Rainflow Cycle Counting


In reality, components are rarely subjected to purely constant amplitude loading, and so for many years various methods for extracting “equivalent” constant amplitude cycles from a random loading
sequence were devised. Methods such as level crossing, range-pair, range-mean, and rainflow all attempt to reduce a random sequence of peaks and valleys to a set of equivalent constant amplitude
cycles.
The concept of an equivalent constant amplitude cycle is perhaps, at first, difficult to understand. On the one hand, the most obvious cycle to be found in any time series will be the one that includes
the maximum and minimum data values. On the other hand, the problem is how to define all the other cycles which are present.

http://www.mscsoftware.com/training_videos/patran/reverb3/Fatigue%20Users%20Guide/... 5/26/2012
Page 4 of 5

Figure 15-5 The Extraction of a Fatigue Cycle


A reversal can be easily understood as a change in the sign of the loading direction. Thus, a constant amplitude signal would consist of two equal and opposite reversals repeated continuously. In the
simplest terms, therefore, a random time series may be considered to be a pair of reversals: the reversal from the maximum to the minimum in the signal and the reversal from minimum to the
maximum in the signal, with all the other reversals effectively “interrupting” these two.
From the point of view of a material, this phenomenon is often referred to as “material memory”, because a material subjected to a sequence of reversing loads, apparently interprets each closed cycle
(matching pair of reversals), as a temporary interruption of a larger strain range, and “remembers” which hysteresis limb applies for this larger event.
Figure 15-5 shows a strain sequence of four turning points, and the stress-strain response of a material to this sequence. The closed hysteresis loop is a cycle, which may be characterized in terms of its
strain range and mean strain. If the stress axis of this diagram is ignored and only successive strain ranges are considered then an algorithm can be developed which will extract cycles from a signal
whatever its units. The rainflow algorithm is able to extract cycles in the way described above, classify them in terms of their range and mean value and store them in a range mean matrix.
The term “rainflow” is derived from an algorithm in which cycles are extracted through a consideration of rain drops flowing down a pagoda roof (the originators of this algorithm were Japanese).
Modern algorithms no longer use this concept although the generic name “rainflow” still persists.

The Rainflow Procedure


The rainflow algorithm used by MSC.Fatigue is based on the standard practice for cycle counting in fatigue analysis as defined by the ASTM designation E 1049-85, (See ASTM standards Vol.
03.01). Note that the form of the algorithm used relies on the fact that the analysis starts at the absolute maximum value in the data set. Under these circumstances, the rainflow cycle count will be
identical to a range pair cycle count which itself starts at the largest value. Differences in the results produced by the two procedures only arise if processing starts at some value other than the absolute
maximum.
It is possible to illustrate the rainflow cycle counting procedure used by considering the simple strain time series shown in Figure 15-6.

Figure 15-6 Illustration of the Rainflow Procedure


The local stress-strain response to the nominal strain time record, shown to the left, reveals that four equivalent constant amplitude cycles are present (i.e., B-C-B, E-F-E, G-H-G and the outside
loop A-D-A). The small event B-C-B is treated as a interruption to the large overall event A-D-A. It is important to note that although, the events B-C-B and G-H-G appear very similar in the
nominal strain record and would be counted as equally damaging by say the range-mean method, the local mean stresses and plastic strains are quite different and so the damage contribution from each
would also be different. It should also be noted that local mean stress cannot be calculated directly from nominal mean strains. The main thrust behind the rainflow cycle counting method, therefore, is
to treat small cycles as interruptions of larger ones.

Local Stress Information


The finite element analysis (FEA) provides a link between applied loads and the stress response at regular locations across the structure. This obviates the need for an engineer to obtain an
approximate stress concentration factor as is normally the case in fatigue analysis. Moreover, the FEA allows the engineer to investigate the fatigue performance for a range of load combinations.

Linear Static Stress Analyses


One of the principles of the MSC.Fatigue package is that the load cases being considered must be analyzed separately unless certain conditions apply (i.e., a stress analysis will be required for each
discrete load applied to the structure). The results from the stress analysis provide a calibration between the applied load and the local stresses across the structure.
Linear elastic stress analysis is used because the fatigue analyzers in MSC.Fatigue carry out various manipulations which include superimposing the FEA stress analysis results and carrying out the
elastic-plastic transformations where necessary. From this, local strain- or stress-time histories at each location representing the strain response due to the simultaneous combination of all the load time
histories are obtained.

Types of Load Cases in FEA


It is important that the units of the loads applied in the FEA correspond to those of the load time histories. To understand the reasoning below, it is necessary to define the following terms:

http://www.mscsoftware.com/training_videos/patran/reverb3/Fatigue%20Users%20Guide/... 5/26/2012
Page 5 of 5

LOAD a value of an applied force in units of force (kN,N, etc).


LOADING a generic term for any external or internal effect which causes straining of a structure.
FEA is not acutely dependent on the actual units of force and length, and hence of stress. The calculation of life in MSC.Fatigue is carried out in S.I. units. However, MSC.Fatigue allows you to work
in a range of unit systems.
Nevertheless, it is still the responsibility of the engineer to ensure the inputs are provided in a consistent manner.
The time history used in a fatigue calculation must be a representation of the time variation in the loading applied in the FEA. For simple cases, this implies a force time history corresponding to a time
variation in the point loading used in the FEA.
There are a number of different kinds of loading possible, each one requiring a different type of time history as shown in Table 15-4.

Table 15-4

Loading Time History Used


Point Loads Load time history at point load application position
Pressure Pressure time history
Temperature Temperature time history
Accelerations Acceleration time history
Distributed Load A related parameter time history1
Moment Moment time history
Scalar Any time history to define a scaled variation of the actual FEA stress-
strain

1
The distributed load is a special case which must be treated by considering a time history of a related quantity such as deflection. In this case you must specify the magnitude of the deflection at a
given location as calculated by the FEA and the corresponding deflection time history.

A load time history will not be required if you wish to define a FE load case as a simple static offset of the stress-strain responses since this does not vary with time though the load case must be
isolated and dealt with in a separate FEA.
Pressures may be applied and pressure time histories used in the fatigue analysis. Pressures and loads may be mixed in a fatigue analysis as long as the units are the same.
Any loads which vary in proportion to and in-phase with each other may be combined together in one FEA. However, the load time history must be a scalar which represents the proportion of these
loads acting on the structure at each increment of time.
Note that, if you plan on using linear elastic results from a transient dynamic or forced vibration FE analysis, none of this applies since you have already defined your time variation in your FE
analysis.

Like

Add New Comment Login

Type your comment here.

Showing 0 comments Sort by popular now

M Subscribe by email S RSS

http://www.mscsoftware.com/training_videos/patran/reverb3/Fatigue%20Users%20Guide/... 5/26/2012

Вам также может понравиться