Sie sind auf Seite 1von 25

HIGH-CYCLE FATIGUE

Historically, most attention has focused on situations that require more than 104
cycles to failure where stress is low and deformation primarily elastic.

The S-N curve

In high-cycle fatigue situations, materials performance is commonly


characterized by an S-N curve, also known as a Whler curve. This is a graph of the
magnitude of a cyclic stress (S) against the logarithmic scale of cycles to failure (N).

S-N curves are derived from tests on samples of the material to be characterized (often
called coupons) where a regular sinusoidal stress is applied by a testing machine
which also counts the number of cycles to failure. This process is sometimes known as
coupon testing. Each coupon test generates a point on the plot though in some cases
there is a run out where the time to failure exceeds that available for the test (see
censoring). Analysis of fatigue data requires techniques from statistics, especially
survival analysis and linear regression.
LOW-CYCLE FATIGUE

Where the stress is high enough for plastic deformation to occur, the account in
terms of stress is less useful and the strain in the material offers a simpler description.
Low-cycle fatigue is usually characterized by the Coffin-Manson relation (published
independently by L. F. Coffin in 1954 and S. S. Manson 1953):

-where:

p /2 is the plastic strain amplitude;


f' is an empirical constant known as the fatigue ductility coefficient, the failure
strain for a single reversal;
2N is the number of reversals to failure (N cycles);
c is an empirical constant known as the fatigue ductility exponent, commonly
ranging from -0.5 to -0.7 for metals in time independent fatigue. Slopes can be
considerably steeper in the presence of creep or environmental interactions.

A similar relationship for materials such as Zirconium, used in the nuclear industry.
Fatigue and fracture mechanics

The account above is purely empirical and, though it allows life prediction and
design assurance, life improvement or design optimisation can be enhanced using
fracture mechanics. It can be developed in four stages.

1. Crack nucleation;
2. Stage I crack-growth;
3. Stage II crack-growth; and
4. Ultimate ductile failure.

Factors that affect fatigue-life

Cyclic stress state: Depending on the complexity of the geometry and the
loading, one or more properties of the stress state need to be considered, such
as stress amplitude, mean stress, biaxiality, in-phase or out-of-phase shear
stress, and load sequence,
Geometry: Notches and variation in cross section throughout a part lead to
stress concentrations where fatigue cracks initiate.
Surface quality. Surface roughness cause microscopic stress concentrations
that lower the fatigue strength. Compressive residual stresses can be introduced
in the surface by e.g. shot peening to increase fatigue life. Such techniques for
producing surface stress are often referred to as peening, whatever the
mechanism used to produce the stress. Low Plasticity Burnishing, Laser
peening, and ultrasonic impact treatment can also produce this surface
compressive stress and can increase the fatigue life of the component. This
improvement is normally observed only for high-cycle fatigue.

Material Type: Fatigue life, as well as the behavior during cyclic loading,
varies widely for different materials, e.g. composites and polymers differ
markedly from metals.

Residual stresses: Welding, cutting, casting, and other manufacturing


processes involving heat or deformation can produce high levels of tensile
residual stress, which decreases the fatigue strength.

Size and distribution of internal defects: Casting defects such as gas


porosity, non-metallic inclusions and shrinkage voids can significantly reduce
fatigue strength.

Direction of loading: For non-isotropic materials, fatigue strength depends on


the direction of the principal stress.
Grain size: For most metals, smaller grains yield longer fatigue lives, however,
the presence of surface defects or scratches will have a greater influence than in
a coarse grained alloy.

Environment: Environmental conditions can cause erosion, corrosion, or gas-


phase embrittlement, which all affect fatigue life. Corrosion fatigue is a
problem encountered in many aggressive environments.

Temperature: Extreme high or low temperatures can decrease fatigue


strength.

Design against fatigue

Dependable design against fatigue-failure requires thorough education and


supervised experience in structural engineering, mechanical engineering, or materials
science. There are three principal approaches to life assurance for mechanical parts
that display increasing degrees of sophistication:

1. Design to keep stress below threshold of fatigue limit (infinite lifetime


concept);
2. Design (conservatively) for a fixed life after which the user is instructed to
replace the part with a new one (a so-called lifed part, finite lifetime concept, or
"safe-life" design practice);
3. Instruct the user to inspect the part periodically for cracks and to replace the
part once a crack exceeds a critical length. This approach usually uses the
technologies of nondestructive testing and requires an accurate prediction of
the rate of crack-growth between inspections. This is often referred to as
damage tolerant design or "retirement-for-cause".

Stopping fatigue

Fatigue cracks that have begun to propagate can sometimes be stopped by


drilling holes, called drill stops, in the path of the fatigue crack.[11] This is not
recommended as a general practice because the hole represents a stress concentration
factor which depends on the size of the hole and geometry, though the hole is typically
less of a stress concentration than the removed tip of the crack. The possibility
remains of a new crack starting in the side of the hole. It is always far better to replace
the cracked part entirely.
Material change

Changes in the materials used in parts can also improve fatigue life. For
example, parts can be made from better fatigue rated metals. Complete replacement
and redesign of parts can also reduce if not eliminate fatigue problems. Thus
helicopter rotor blades and propellers in metal are being replaced by composite
equivalents. They are not only lighter, but also much more resistant to fatigue. They
are more expensive, but the extra cost is amply repaid by their greater integrity, since
loss of a rotor blade usually leads to total loss of the aircraft. A similar argument has
been made for replacement of metal fuselages, wings and tails of aircraft.

Probabilistic nature of fatigue

As coupons sampled from a homogeneous frame will manifest variation in


their number of cycles to failure, the S-N curve should more properly be an S-N-P
curve capturing the probability of failure after a given number of cycles of a certain
stress. Probability distributions that are common in data analysis and in design against
fatigue include the lognormal distribution, extreme value distribution, Birnbaum
Saunders distribution, and Weibull distribution.

Complex loadings

Spectrum loading

In practice, a mechanical part is exposed to a complex, often random, sequence of


loads, large and small. In order to assess the safe life of such a part:

1. Reduce the complex loading to a series of simple cyclic loadings using a


technique such as rainflow analysis;
2. Create a histogram of cyclic stress from the rainflow analysis to form a fatigue
damage spectrum;
3. For each stress level, calculate the degree of cumulative damage incurred from
the S-N curve; and
4. Combine the individual contributions using an algorithm such as Miner's rule.

Miner's rule

In 1945, M. A. Miner popularised a rule that had first been proposed by A.


Palmgren in 1924. The rule, variously called Miner's rule or the Palmgren-Miner
linear damage hypothesis, states that where there are k different stress magnitudes in a
spectrum, S i (1 i k), each contributing ni(Si) cycles, then if Ni(Si) is the number of
cycles to failure of a constant stress reversal Si, failure occurs when:

C is experimentally found to be between 0.7 and 2.2. Usually for design purposes, C is
assumed to be 1.

This can be thought of as assessing what proportion of life is consumed by stress


reversal at each magnitude then forming a linear combination of their aggregate.

Though Miner's rule is a useful approximation in many circumstances, it has several


major limitations:

1. It fails to recognise the probabilistic nature of fatigue and there is no simple


way to relate life predicted by the rule with the characteristics of a probability
distribution. Industry analysts often use design curves, adjusted to account for
scatter, to calculate Ni(S i).
2. There is sometimes an effect in the order in which the reversals occur. In some
circumstances, cycles of low stress followed by high stress cause more damage
than would be predicted by the rule. It does not consider the effect of overload
or high stress which may result in a compressive residual stress. High stress
followed by low stress may have less damage due to the presence of
compressive residual stress.
Paris' Relationship

Typical fatigue crack growth rate graph.

In Fracture mechanics, Anderson, Gomez and Paris derived relationships for


the stage II crack growth with cycles N, in terms of the cyclical component K of the
Stress Intensity Factor K[8]

where a is the crack length and m is typically in the range 3 to 5 (for metals).

This relationship was later modified (by Forman, 1967 [9]) to make better allowance for
the mean stress, by introducing a factor depending on (1-R) where R = min stress/max
stress, in the denominator.

Miners law:

The effects of fatigue are cumulative and it is difficult to predict the fatigue life
of a component that works under varying conditions of stress. For example an aircraft
in a strong weather conditions show lesser fatigue life. Hence it is important to
calculate the fatigue life by finite endurance limit.
According to Miners law the total life of a part can be estimated by adding up
the percentage of life consumed by each over stress cycle. If an n1 cycle in a cyclic
loading leads to failure after N1 cycles then n1/N1 is the proportion of damage that
has occurred. If the stress amplitude change to n2 cycles and at that amplitude failure
occurs after N2 cycles then n2/N2 is a measure of the proportion of damage caused
during that period. According to Miners law the component will fail when the sum of
the entire cyclic ration equals to unity.

Miners law is also called as Fatigue Miners rule or cumulative damage rule.

Bauschinger effect:

According Bauschinger under cyclic stress (fatigue) the proportionality limit


(yield strength) of the material does not remain constant but varies according to the
direction ofstress. During plastic deformation, the yield strength of the metal increases
in the direction of plastic low when loaded beyond the elastic limit. But the plastic
deformation would start at a lower yield stress is applied in the direction opposite to
the initial direction. This is a due to the fact that under the reverse stress caused by
initial deformation increases the stresses. This phenomenon is called as Bauschinger
effect of elastic hysteresis.
In graph (a) the points A and B on the curve represents the yield stresses of the
metal when it is loaded in tension and compression respectively. The yield stresses at
A and B will be equal in magnitude and opposite in sign. Supposing if the tensile load
is gradually applied at higher stresses than the yield stress (i.e., stressed beyond elastic
limit) we obtain the curve OF in graph (b). The pint F is at a higher stress than the
yield point A. Now if the load is gradually removed then the curve will follow the
path FE instead of FO. This means that the metal has developed a permanent strain
OE. If the metal is reloaded in a direction opposite to the original tip direction (i.e.,
compressive load is applied gradually), the plastic flow begins at D. The compressive
stress at D is low than the original value of compressive yield stress at F. This
decrease in compressive yield stress after the tensile loading of the metal is known as
Bauschinger effect.

Similarly if the metal is loaded initially in compression, the Bauschinger effect


will be observed when the metal is reloaded in tension at point G. Thus the reduction
in compressive or tensile stress due to the presence of the presence of residual stress
even after the removal of load. These residual stress cause the dislocation to move
more easily in a direction opposite to the original direction even at low stresses, also
when the slip direction is reversed, dislocation of opposite signs may be created which
attract and cause slip easier. The total result is the softening of the metal. Hence the
plastic flow begins at a lower stress and failure by fatigue occurs by repeated by cyclic
loading.

Strain hardening

Strain hardening or cold working, is the strengthening of a metal by plastic


deformation. This strengthening occurs because of dislocation movements within the
crystal structure of the material. Any material with a reasonably high melting point
such as metals and alloys can be strengthened in this fashion. Alloys not amenable to
heat treatment, including low-carbon steel, are often work-hardened. Some materials
cannot be work-hardened at normal ambient temperatures, such as indium, however
others can only be strengthened via work hardening, such as pure copper and
aluminum.

Work hardening may be desirable or undesirable depending on the context. An


example of undesirable work hardening is during machining when early passes of a
cutter inadvertently work-harden the workpiece surface, causing damage to the cutter
during the later passes. An example of desirable work hardening is that which occurs
in metalworking processes that intentionally induce plastic deformation to exact a
shape change. These processes are known as cold working or cold forming
processes. They are characterized by shaping the workpiece at a temperature below its
recrystallization temperature, usually at the ambient temperature. Cold forming
techniques are usually classified into four major groups: squeezing, bending, drawing,
and shearing. Examples of applications include the heading of bolts and cap screws
and the finishing of cold rolled steel.

Theory

Before work hardening, the lattice of the material exhibits a regular, nearly
defect-free pattern (almost no dislocations). The defect-free lattice can be created or
restored at any time by annealing. As the material is work hardened it becomes
increasingly saturated with new dislocations, and more dislocations are prevented
from nucleating (a resistance to dislocation-formation develops). This resistance to
dislocation-formation manifests itself as a resistance to plastic deformation; hence, the
observed strengthening.

In metallic crystals, irreversible deformation is usually carried out on a


microscopic scale by defects called dislocations, which are created by fluctuations in
local stress fields within the material culminating in a lattice rearrangement as the
dislocations propagate through the lattice. At normal temperatures the dislocations are
not annihilated by annealing. Instead, the dislocations accumulate, interact with one
another, and serve as pinning points or obstacles that significantly impede their
motion. This leads to an increase in the yield strength of the material and a subsequent
decrease in ductility.

Such deformation increases the concentration of dislocations which may


subsequently form low-angle grain boundaries surrounding sub-grains. Cold working
generally results in higher yield strength as a result of the increased number of
dislocations and the Hall-Petch effect of the sub-grains, and a decrease in ductility.
The effects of cold working may be reversed by annealing the material at high
temperatures where recovery and recrystallization reduce the dislocation density.

A material's work hardenability can be predicted by analyzing a stress-strain curve, or


studied in context by performing hardness tests before and after a process. [citation needed]

Elastic and plastic deformation

Work hardening is a consequence of plastic deformation, a permanent change


in shape. This is distinct from elastic deformation, which is reversible. Most materials
do not exhibit only one or the other, but rather a combination of the two. The
following discussion mostly applies to metals, especially steels, which are well
studied. Work hardening occurs most notably for ductile materials such as metals.
Ductility is the ability of a material to undergo large plastic deformations before
fracture (for example, bending a steel rod until it finally breaks).

The tensile test is widely used to study deformation mechanisms. This is


because under compression, most materials will experience trivial (lattice mismatch)
and non-trivial (buckling) events before plastic deformation or fracture occur. Hence
the intermediate processes that occur to the material under uniaxial compression
before the incidence of plastic deformation make the compressive test fraught with
difficulties.

A material generally deforms elastically if it is under the influence of small


forces, allowing the material to readily return to its original shape when the deforming
force is removed. This phenomenon is called elastic deformation. This behavior in
materials is described by Hooke's Law. Materials behave elastically until the
deforming force increases beyond the elastic limit, also known as the yield stress. At
this point, the material is rendered permanently deformed and fails to return to its
original shape when the force is removed. This phenomenon is called plastic
deformation. For example, if one stretches a coil spring up to a certain point, it will
return to its original shape, but once it is stretched beyond the elastic limit, it will
remain deformed and won't return to its original state.

Elastic deformation stretches atomic bonds in the material away from their
equilibrium radius of separation of a bond, without applying enough energy to break
the inter-atomic bonds. Plastic deformation, on the other hand, breaks inter-atomic
bonds, and involves the rearrangement of atoms in a solid material.

Dislocations and lattice strain fields

In materials science parlance, dislocations are defined as line defects in a


material's crystal structure. They are surrounded by relatively strained (and weaker)
bonds than the bonds between the constituents of the regular crystal lattice. This
explains why these bonds break first during plastic deformation. Like any
thermodynamic system, the crystals tend to lower their energy through bond formation
between constituents of the crystal. Thus the dislocations interact with one another
and the atoms of the crystal. This results in a lower but energetically favorable energy
conformation of the crystal. Dislocations are a "negative-entity" in that they do not
exist: they are merely vacancies in the host medium which does exist. As such, the
material itself does not move much. To a much greater extent visible "motion" is
movement in a bonding pattern of largely stationary atoms.
The strained bonds around a dislocation are characterized by lattice strain
fields. For example, there are compressively strained bonds directly next to an edge
dislocation and tensilely strained bonds beyond the end of an edge dislocation. These
form compressive strain fields and tensile strain fields, respectively. Strain fields are
analogous to electric fields in certain ways. Additionally, the strain fields of
dislocations obey the laws of attraction and repulsion.

The visible (macroscopic) results of plastic deformation are the result of


microscopic dislocation motion. For example, the stretching of a steel rod in a tensile
tester is accommodated through dislocation motion on the atomic scale.

Increase of dislocations and work hardening

Figure 1: The yield stress of an ordered material has a half-root dependency on the
number of dislocations present.

Increase in the number of dislocations is a quantification of work hardening.


Plastic deformation occurs as a consequence of work being done on a material; energy
is added to the material. In addition, the energy is almost always applied fast enough
and in large enough magnitude to not only move existing dislocations, but also to
produce a great number of new dislocations by jarring or working the material
sufficiently enough. New dislocations are generated in proximity to a Frank-Read
source.

Yield strength is increased in a cold-worked material. Using lattice strain


fields, it can be shown that an environment filled with dislocations will hinder the
movement of any one dislocation. Because dislocation motion is hindered, plastic
deformation cannot occur at normal stresses. Upon application of stresses just beyond
the yield strength of the non-cold-worked material, a cold-worked material will
continue to deform using the only mechanism available: elastic deformation. The
regular scheme of stretching or compressing of electrical bonds (without dislocation
motion) continues to occur, and the modulus of elasticity is unchanged. Eventually the
stress is great enough to overcome the strain-field interactions and plastic deformation
resumes.

However, ductility of a work-hardened material is decreased. Ductility is the


extent to which a material can undergo plastic deformation, that is, it is how far a
material can be plastically deformed before fracture. A cold-worked material is, in
effect, a normal material that has already been extended through part of its allowed
plastic deformation. If dislocation motion and plastic deformation have been hindered
enough by dislocation accumulation, and stretching of electronic bonds and elastic
deformation have reached their limit, a third mode of deformation occurs: fracture.

Quantification of work hardening

The stress, , of dislocation is dependent on the shear modulus, G, the lattice


constant, b, and the dislocation density, :

where 0 is the intrinsic strength of the material with low dislocation density and is a
correction factor specific to the material.

As shown in Figure 1 and the equation above, work hardening has a half root
dependency on the number of dislocations. The material exhibits high strength if there
are either high levels of dislocations (greater than 10 14 dislocations per m2) or no
dislocations. A moderate number of dislocations (between 107 and 109 dislocations per
m2) typically results in low strength.

Example

For an extreme example, in a tensile test a bar of steel is strained to just before
the distance at which it usually fractures. The load is released smoothly and the
material relieves some of its strain by decreasing in length. The decrease in length is
called the elastic recovery, and the end result is a work-hardened steel bar. The
fraction of length recovered (length recovered/original length) is equal to the yield-
stress divided by the modulus of elasticity. (Here we discuss true stress in order to
account for the drastic decrease in diameter in this tensile test.) The length recovered
after removing a load from a material just before it breaks is equal to the length
recovered after removing a load just before it enters plastic deformation.
The work-hardened steel bar has a large enough number of dislocations that the
strain field interaction prevents all plastic deformation. Subsequent deformation
requires a stress that varies linearly with the strain observed; the slope of the graph of
stress vs. strain is the modulus of elasticity, as usual.

The work-hardened steel bar fractures when the applied stress exceeds the
usual fracture stress and the strain exceeds usual fracture strain. This may be
considered to be the elastic limit and the yield stress is now equal to the fracture
toughness, which is of course, much higher than a non-work-hardened-steel yield
stress.

The amount of plastic deformation possible is zero, which is obviously less


than the amount of plastic deformation possible for a non-work-hardened material.
Thus, the ductility of the cold-worked bar is reduced.

Additionally, jewelers will construct structurally sound rings and other


wearable objects (especially those worn on the hands) that require much more
durability (than earrings for example) by utilizing a material's ability to be work
hardened. While casting rings is done for a number of economical reasons (saving a
great deal of time and cost of labor), a master jeweler may utilize the ability of a
material to be work hardened and apply some combination of cold forming techniques
during the production of a piece.

Empirical relations

There are two common mathematical descriptions of the work hardening


phenomenon. Hollomon's equation is a power law relationship between the stress and
the amount of plastic strain:

where is the stress, K is the strength index, p is the plastic strain and n is the strain
hardening exponent. Ludwik's equation is similar but includes the yield stress:

If a material has been subjected to prior deformation (at low temperature) then
the yield stress will be increased by a factor depending on the amount of prior plastic
strain 0:
The constant K is structure dependent and is influenced by processing while n
is a material property normally lying in the range 0.20.5. The strain hardening index
can be described by:

This equation can be evaluated from the slope of a log () - log () plot.
Rearranging allows a determination of the rate of strain hardening at a given stress and
strain:

Cyclic Stress-Strain Graphs

Cyclic Hardening
Stresses increase with increasing number of cycles
Cyclic Softening
Stresses decrease with increasing number of cycles
Cycle counting techniques

Rainflow Cycle Counting Method

Counting methods have initially been developed for the study of fatigue damage
generated in aeronautical structures. Since different results have been obtained from
different methods, errors could be taken in the calculations for some of them. Level
crossing counting, peak counting, simple range counting and rainflow counting are the
methods which are using stress or deformation ranges. One of the preferred methods is
the rainflow counting method.
Rainflow cycle counting method has initially been proposed by M.Matsuiski
and T.Endo to count the cycles or the half cycles of strain-time signals. [14] Counting
is carried out on the basis of the stress-strain behavior of the material. This is
illustrated in Figure 3.1. As the material deforms from point a to b, it follows a path
described by the cyclic stress-strain curve. At point b, the load is reversed and the
material elastically unloads to point c. When the load is reapplied from c to d, the
material elastically deforms to point b, where the material remembers its prior history,
i.e. from a to b, and deformation continues along path a to d as if event b-c never
occurred.
The signal measured, in general, a random stress S(t) is not only made up of a
peak alone between two passages by zero, but also several peaks appear, which makes
difficult the determination of the number of cycles absorbed by the structure. An
example for the random stress data is shown in Figure 3.2.

Figure 3.2. Random stress fluctuation

The counting of peaks makes it possible to constitute a histogram of the peaks


of the random stress which can then be transformed into a stress spectrum giving the
number of events for lower than a given stress value. The stress spectrum is thus a
representation of the statistical distribution of the characteristic amplitudes of the
random stress as a function of time. The origin of the name of rainflow counting
method which is called Pagoda Roof Method can be explained as that the time axis
is vertical and the random stress S(t) represents a series of roofs on which water falls.
The rules of the flow can be shown as in Figure 3.3.

Figure 3.3. The drop released from the largest peak


The origin of the random stress is placed on the axis at the abscissa of the largest peak
of the random stress. Water drops are sequentially released at each extreme. It can be
agreed that the tops of the roofs are on the right of the axis, bottoms of the roofs are on
the left.
If the fall starts from a peak:
a) the drop will stop if it meets an opposing peak larger than that of departure,
b) it will also stop if it meets the path traversed by another drop, previously
determined as shown in Figure 3.4,
c) the drop can fall on another roof and to continue to slip according to rules a and b.

Figure 3.5. Drop departure from a valley

Figure 3.6. Flow rule of the drop from a valley

As the fundamentals of the original definition of the rainflow cycle counting given
above, the cycles are identified in a random variable amplitude loading sequence in
Figure 3.7 as an example. First, the stress S(t) is transformed to a process of peaks and
valleys. Then the time axis is rotated so that it points downward. At both peaks and
valleys, water sources are considered. Water flows downward according to the
following rules:
1. A rainflow path starting at a valley will continue down the pagoda roofs, until it
encounters a valley that is more negative than the origin. From the figure, the path that
starts at A will end at E.
2. A rainflow path is terminated when it encounters flow from a previous path. For
example, the path that starts at C is terminated as shown.
3. A new path is not started until the path under consideration is stopped.
4. Valley-generated half-cycles are defined for the entire record. For each cycle, the
stress range Si is the vertical excursion of a path. The mean Si is the midpoint.
5. The process is repeated in reverse with peak-generated rainflow paths. For a
sufficiently long record, each valley-generated half-cycle will match a peak-generated
half-cycle to form a whole cycle.

Figure 3.7. Rainflow cycle counting


Practical definition of rainflow cycle counting

The results obtained from Figure 3.8 are tabulated in Table 3.1. It gives the number of
cycle counts in the specific events.
Table 3.1. Cycle counts

Various methods of counting were proposed, leading to different results and, thus, for
some, to errors in the calculation of the fatigue lives. Although various methods may
still be in use, Rainflow Counting is the preferred method. This method includes a
family of various computer algorithms. Older methods which often utilized analog
logic circuits are
Level Crossing,
Peak Counting,
Simple Range.
LEVEL CROSSING

PEAK COUNTING
SIMPLE RANGE

Cumulative damage analysis:

The fatigue life prediction process or cumulative damage analysis for a critical
region in a component or structure consists of several closely interrelated steps as can
be seen in Figure 2.1, separately. A combination of the load history (Service Loads),
stress concentration factors (Stress Analysis) and cyclic stress-strain properties of the
materials (Material Properties) can be used to simulate the local uniaxial stress-strain
response in critical areas. Through this process it is possible to develop good estimates
of local stress amplitudes, mean stresses and elastic and plastic strain components for
each excursion in the load history. Rainflow counting can be used to identify local
cyclic events in a manner consistent with the basic material behavior. The damage
contribution of these events is calculated by comparison with material fatigue data
generated in laboratory tests on small specimens. The damage fractions are summed
linearly to give an estimate of the total damage for a particular load history.
Four Alternative Approaches

The general concept of damage that would or could lead to failure was rather vague
and nebulous until Miner gave it a specific form and it became recognized. With
background from Palmgren, Miner [3] postulated the damage form for fatigue as

where ni is the number of applied cycles at nominal stress i and Ni is the limit
number of cycles to failure at the same stress and for the same cycle type. Thus each
value ni/Ni is viewed as a quanta of damage, the sum of which specifies failure. As
with all cumulative damage forms, when the left hand side of (1) is less than one, it
still quantifies the damage level but does not imply failure. The spectrum of values of
N( ) versus constant stress, , is as shown in Fig. 1. Relation (1) then allows the life
prediction for a combination of different load levels. All of the fatigue conditions
considered here will be taken to be of the same frequency and cycle type.

Das könnte Ihnen auch gefallen