You are on page 1of 13

Fatigue

Fatigue is a type of fracture that occurs in materials that are subjected to changing or varying stresses over time. Fatigue is major problem in many of the modern machines and devices we use today, such as bridges and airplanes. Fatigue is mainly caused by the environment in which the materials are utilized. Therefore, much research has been done in the field of fatigue fracture resistance to counter the effect of varying stress on the material. Even recently, April 1st 2011, Southwest Flight 812 suffered from a catastrophic fatigue fracture which happened 18 minutes into its flight[3]. The fuselage skin ripped open causing a massive loss in cabin pressure and was forced to land on a remote military base in Yuma, Arizona[5]. The cause of the damage isnt always apparent due to the nature of fatigue fracture often occurring at low stresses and elevated temperatures in most materials. It is accidents like this one that have been a leading motivation in fatigue research.

Contents
1. Fatigue Fracture/Cyclic Stress-Strain (Brian) 2. High and Low Cycle Fatigue (Dan) 2.1 High Cycle Fatigue 2.2 Low Cycle Fatigue 3. Fatigue Resistance and Designing Against Fatigue Fracture (Dan) 3.1 Fatigue Resistance 3.2 Design Against Fatigue 4. Fatigue Fracture and Crack Growth Rates (Brian) 5. Types of Fatigue (Alex) 5.1 Fatigue in Metals and Polymers 5.2 Fatigue in Composites 6. Creep Fatigue (Alex) 7. See Also 8. References

Fatigue Fracture/Cyclic Stress-Strain


Fatigue fracture, like yielding, is due to formations of cracks at the microscopic level and lengthened by continued applications of stress. It differs however in the manner the stress is applied. Fatigue fracture is instigated by cyclical stresses on the material. This is achieved through a process of crack nucleation and pre-existent cracks in a material[1]. The stress instigated during cyclical stress is mostly sinusoidal in nature, that is varying with a specific frequency of stress application[2]. Each peak is denoted as either the max stress max or min which are equal and opposite in sign (tension and compression respectively). The applied stress can be torsional, tensile, or flexural in nature. The following figure illustrates each type of stress.

Figure 1[3] Torsional, tensile, and flexural applied stresses The relationship between stress and cycles to failure are related using a S-N plot. The plot shows the applied max stress versus the amount of cycles it took for the material to fail, otherwise known as the fatigue life of a material.

Figure 2[4] S-N plot showing a material that has an endurance limit (A) and one which does not have an endurance limit (B). The above two curves exhibit two different kind of fatigue in materials. Curve A denotes a material that has a fatigue limit. A fatigue limit (also called an endurance limit) denotes the critical stress below which fatigue fracture will not happen. Curve B denotes a fatigue curve that does not have a fatigue limit. It does however have fatigue strengths. That is to say for a given stress their is a given fatigue life or number of cycles to failure of that material. The maximum number of cycles it takes for a material to fracture is called fatigue strength. Fatigue fracture is a ductile fracture, and therefore occurs by non-uniform plastic deformation. Microcracks and voids form after a certain number of cycles and grow proportional to the number of cycles eventually grow large enough to overcome recovery mechanisms and move quickly to fracture. The rate of this crack nucleation is proportional to the frequency of the applied stress. Fatigue fracture is mostly centered on metals. Fatigue fracture is easily identifiable by outwardly arcing lines of deformation from the crack initiation point, which indicates gradual crack growth rather than sudden growth. Also, the slow

crack growth creates fatigue striations whose displacement from one another is proportional to crack growth per cycle of stress.

High and Low Cycle Fatigue

High-cycle Fatigue
High-cycle fatigue is when the number of cycles to failure is large, typically when the number of cycles to failure, Nf, is greater than 103.

Low-cycle Fatigue
Low- cycle fatigue is when the number of cycles to failure is small, typically when the number of cycles to failure, Nf, is less than 103.

Fatigue Resistance and Designing Against Fatigue Fracture

Fatigue Resistance
In order for fatigue fracture to occur some portion of a time-varying stress must be tensile in nature; fatigue fracture will not occur in compression and as result failure of engineering structures by fatigue is usually under stochastic, or random, loading. The analysis of a materials fatigue resistance usually involves implementing of a well characterized stress/strain cycle despite the fact that the stress cycle the material would experience in practice is in fact not well characterized and most likely random. An example of this would be the stress/strain conditions an aircraft wing would experience during flight would be related to the weather, turbulence, other flight conditions, etc., and therefore random. A materials fatigue resistance is also affected by the frequency of the applied stress. For polymeric materials the effect of frequency can be seen at relatively low frequencies whereas for metals and metal alloys the fatigue behavior only depends on cyclical frequency at relatively high frequencies. The fatigue behavior of metals can be considered independent of frequency for frequencies below about 200 Hz. Since this frequency range encompasses the majority of engineering applications, frequency ranges above 200 Hz are often not considered.

Figure 3[1] Stress vs. time plots for (a) a material undergoing random loading superimposed on a positive mean stress (b) a material subject to a rotating beam stress test where the load is both compressive and tensile and (c) a material undergoing a cyclical stress test where time-varying stress is superimposed on a constant mean stress. Because of the important role of plasticity in metal fatigue, it is fundamentally more sound to assess a materials fatigue response under conditions of a specified cyclical strain, rather than stress[1]. Despite this, the traditional testing procedure to determine a materials fatigue resistance is to perform a series of test where the stress amplitude is varied while maintaining a fixed mean stress and measure the number of stress cycles,Nf, required to cause fatigue fracture. For some steels, for example, there is a stress amplitude below which fracture will not take place no matter how many cycles the material is subject to. This point is called the fatigue limit or endurance limit of the steel. As a rule of thumb the fatigue limit is approximately 40% of the tensile strength of the steel when the limit is obtained for mean = 0. For some metals, such as many aluminum alloys, a fatigue limit does not exist and the material will always fracture by fatigue after some sufficiently large number of stress cycles. For materials not exhibiting a fatigue limit, an effective endurance limit is used in design for fatigue resistance. For materials with a fatigue limit, the fatigue limit is commonly used as a fatigue design parameter. Several empirical engineering estimates have been developed to relate the endurance limit to the mean stress. One of these is the Goodman equation which is a generally considered to be a conservative approximation. The Goodman equation is as follows:

(Equation 1)

Where fat is the endurance limit when mean = 0 and T.S. is the materials tensile stress[1]. When the number of cycles to failure is large, referred to as high-cycle fatigue (typically greater than 103 cycles), the material undergoes plastic deformation due to the macroscopic stress level. As result, the elastic strain range, el, is coupled to the stress range by el = /E where E is the Youngs modulus. When the material undergoes low-cycle fatigue it is subject to both macroscopic and microscopic plastic strain and the plastic strain range is much greater than the elastic so that pl where pl is the plastic strain range and is related to via the materials cyclic hardening response and is the total strain range. In general the total strain range is the sum of the plastic and elastic strain ranges where plastic strain range dominates in low-cycle fatigue and elastic strain range dominates in high-cycle fatigue. The

transition from low-cycle fatigue to high-cycle fatigue typically takes place at value of 103 cycles where

. For high-cycle fatigue the elastic strain range is correlated to Nf through the equation

(Equation 2)

Where E is the modulus, and b are determined by curve fitting, and has been found to scale with the materials tensile strength[1]. From this equation it is apparent that a strong material is best for applications where the material will experience high-cycle fatigue.

Figure 4[1] strain range vs number of cycles for some engineering materials. Note that the hardened and annealed 4340 exhibit a fatigue limit while the copper does not. A majority of the time a material spends in high-cycle fatigue is spent nucleating cracks, caused by local plastic deformation, often on the surface of the material. As result, increasing the materials surface strength can delay crack nucleation. Cold working (shot peening) and specialized chemical and thermal surface treatments which increase the surface yield strength (e.g. carburizing, nitriding, and surface martensite formation in steels). For low-cycle fatigue some empirical relationships between pl and Nf have also been developed. One of these is

(Equation 3) is close to the tensile ductility[1].

Where c is 0.5-0.7 typically and

The general relationship between and Nf is the sum of the strain amplitude-cycles to failure relationship at the extremes of high-cycle and low-cycle fatigue. The equation for this general relationship is given below[1].

(Equation 4)

Figure 5[1] Strain amplitude vs. number of stress/strain reversals. At high cycles the slope of the line is -b and at low Nf values the slope is -c. The two lines then intersect at approximately .

Design Against Fatigue


In designing against fatigue there are three approaches: (1) choosing a material which will experience a stress below its fatigue limit (with a safety factor included) (2) if the material has no fatigue limit a number of cycles before replacement would be given by the manufacturer and after this set number of cycles the part(s) would be replaced and (3) periodically assessing the part(s) for cracks and damage and repairing or replacing as needed.

Fatigue Fracture and Crack Growth Rates


Stage II crack growth plays a large role in fatigue fracture due to the slow crack mechanism engaged in fatigue fracture. Stage II crack growth is most prevalent in low cycle fracture. Preexisting cracks and nucleated cracks both play strong roles in fracture. Preexisting cracks in a material can be used to

determine fatigue strength/life based on geometry of the crack from basic fracture mechanics. Crack rates are approximated by the following relation:

where

(Equation 5)

is the applied stress and c is the crack length. A is a material constant and m is an empirically derived constant. Crack rates are given as the change in crack length per change in cycles of fatigue. Also the crack rate increases with increasing stress ratio R which is min stress divided by the max stress and thusly fracture sooner. Also there is a critical value of = at which cracks propagate above , then tensile failure and crack advance occur and dont propagate below. As K reaches simultaneously causing the material to fracture[1].

Figure 6[1] Plot of Crack length vs. Number of cycles. High values lead to large crack rates, which lead to higher crack advance rates which finally leads to

are associated with tensile fracture. High values are also associated with low-cycle fatigue. Low closely spaced striations which are not very apparent. Medium values are where characteristic fatigue striations become apparent.

Figure 7[1] Thusly stage II crack growth and the number of striations in the material are directly proportional to the crack growth rate and relationship. is thusly the critical crack length that will take the material to failure and being the original crack length. The number of striations is related to the critical crack length by the above equation derivation ending with the final relation below:

( Equation 6)

is a material constant. When stage II crack growth occurs correlates to . Fatigue fracture is not very micro-structurally dependent, by the materials response to applied cyclical stresses. Crack rates share a very strong relation to elastic modulus of a material (illustrated in the following figure).

Figure 8[1]

Types of Fatigue

Fatigue in Metals and Polymers


Metal and Polymeric fatigue is the progressive and localized structural damage that occurs when a material is subjected to cyclic loadings. Today, metal fatigue is very often associated with polymeric fatigue, even though they are not the same materials at all. Polymeric use primarily substitutes for metals, and through the years of developing new ways to make polymers stronger, its sustainability has increased exponentially. Even though polymers can be very durable and resistant to stress or strain, they do at some point fatigue similar to metals. This is due to the exposure of cyclical stress or strain. Metals and polymers are already different in the fact that polymers are viscoelastic and show hysteretic elastic effects commonly. On the other hand this is not the case for metals, because they tend to only have linear elastic behavior. Yet the relationship between stress or strain amplitude and fatigue life are asserted for polymers in the same way as for metals. Some polymeric materials also exhibit the same endurance levels as metals. In the figure below, the a - Nf relationship for polymers shown schematically:

Figure 9[1]

In Region I the mechanical characteristics of the material are shown. If the maximum stress during the first cycle of a fatigue test exceeds the materials craze yield strength, crazes are formed and serve as crack nuclei.

In Region II the slope of the stress amplitude-cycles to failure are shown. Also crack propagation rates become increasingly important in determining material lifetime as strain or stress amplitude is increased. This fatigue is also common in metals. At higher stress amplitudes of this Region, crack propagation is frequently associated with crazing. In Region III the high-cycle fatigue region is shown. It is characterized by a fairly negative slope of the a - Nf. At the low stress and strain amplitudes characteristic to Region III, fatigue cracks nucleate by heterogeneous, microscopic flow processes. Here also at high temperatures at which fatigue controls polymers cyclic behavior, crazing is very frequently the flow that is associated with high-cyclic fatiguecrack nucleation. In general, these three regions of the graph all have a tendency to arrive at the practice of crazing.

Fatigue in Composites
Cyclic damage accumulates in the material, by forming fatigue cracks in the matrix. Mostly caused by fiber oxidation and corrosion. Even though there is not much known and written about fatigue in composites, it is still very important to explore. Usually a common composite is used to illustrate the fatigue. This composite is an aligned fiber whose matrix is an epoxy resin. Cyclic damage accumulates in the material by forming fatigue cracks in the matrix. The fibers in the wake of the crack experience fatigue damage, even though the matrices carry small fractions of the applied stress. As the crack wake damage increases, composite properties degrade. An example for this is when the composite modulus decreases, this leads up to composite fracture. Also during cyclical loading this constant composite weakening constitutes engineering failure. This failure under cyclical loading is defined by a different criterion than applies to other material classes.

Creep Fatigue
In the initial stage, or primary creep, the strain rate is relatively high, but slows with increasing strain. This is due to work hardening. The strain rate eventually reaches a minimum and becomes near constant. This is due to the balance between work hardening and annealing (thermal softening). This stage is known as secondary or steady-state creep. This stage is the most understood. The characterized "creep strain rate" typically refers to the rate in this secondary stage. Stress dependence of this rate depends on the creep mechanism. In tertiary creep, the strain rate exponentially increases with strain because of necking phenomena. Creep fatigue is understood to be an environmentally induced fracture that is considerably important when this is one of the sole factors responsible for material failure. Creep fatigue is mostly explored in high-temperature applications because it is there where the importance of it comes to show. As well as most fatigue environments are considered in high-temperature circumstances. Shown below is the demonstration of how cyclical strain may accelerate void growth in a creeping solid:

Figure 10[1] Illustration showing: a) A cavity situated on a grain boundary. b) During boundary sliding the respective halves are displaced. c) A diffusive flux that attempts to maintain equilibrium dihedral angle at the boundary-cavity junction results. d) Cavity growth is caused. e) Process is repeated on stress reversal.

The effect of temperature on the plastic strain amplitude-cycles to failure relationship is shown in figure 11 (a) below. Notice at very high strain amplitudes temperature generally reduces a materials fatigue life. This graph constructively shows a fixed cycle loading frequency. In figure 11 (b) a schematic illustration of the effect of cyclical frequency and environment of fatigue life is represented. Air environments decrease fatigue life.

Figure 11[1] Creep fatigue interactions in metals and polymers i greatly acknowledged by the previous information. It is very clear that temperature plays a very important role in metal fatigue at which diffusion becomes significant. Temperatures play even a larger role at room-temperature for fatigue in polymers, because that means it is usually a relatively high homologous temperature for this material class. The viscoelastic component of deformation in polymers also provides for significant hysteretic stress-strain behavior even without permanent deformation. Creep fatigue must become minimized in polymers since they are becoming increasingly used as substitutes for metals.

See Also
Aloha Airlines Flight 243 Fracture Fatigue Failure Analysis: Analyzing Material Fatigue

References
1. Courtney, T.H. (2005). Mechanical behavior of materials. Long Grove, IL: Waveland Press, Inc. 2. Callister, Jr., W.D. (2008). Fundamentals of materials science and engineering: an integrated approach. Hoboken NJ: Waveland Press, Inc. 3. http://www.sv.vt.edu/classes/MSE2094_NoteBook/97ClassProj/anal/kelly/fatigue.html 4. http://www.fea-optimization.com/ETBX/stresslife_help.html 5. http://articles.latimes.com/2011/apr/04/nation/la-na-southwest-20110404

You might also like