Anda di halaman 1dari 4

Material

E E A A
ACCELERATED FATIGUE TESTING

S S

E E

16 15
A MPTIAC
Brigitte Battat AMPTIAC, Rome, NY

This article is the second installment of excerpts from a new AMPTIAC State-of-the-Art Review entitled Accelerated Testing: A Methodology for Determining Life and Degradation of Materials that will be published in the near future. The first installment can be found in Volume 5, Number 1 of the AMPTIAC Newsletter. The objective of the present article is to define fatigue and to succinctly describe the mechanisms governing these types of failures.

Why Accelerated Testing? The competitive climate of technology and industry today is driving the product development cycle to shorter time-scales with fewer resources. The roles of materials science and process analysis in the development cycle cannot be overemphasized, and when carried out properly, they save time and resources, contributing to better and more reliable products. Accelerated testing consists of test methods that deliberately shorten (in a measured way) the life of the tested product or accelerate the degradation of the products performance. Most accelerated testing is carried out on materials or products to characterize their degradation mechanisms (e.g., fatigue, creep, cracking, wear, corrosion / oxidation, weathering). Accelerated testing uses models and analytical representations that describe the mechanisms governing the process investigated. Mechanistic models represent the association between physical processes and critical parameters affecting the life of the material or the component. These models can be classified into three broad categories: Activation Models Inverse Power Models Materials Models These classifications recognize that there are three driving factors in general, namely temperature, loading and material properties. For example, fatigue failure takes place, when either loads (stresses) are too high, or material toughness is too low. Consequently, low material toughness may be due to poor inherent strength or excessively high operating temperature. The Phenomenon of Fatigue[1] To address designers concerns about a materials ability to withstand cyclic loading, cyclic fatigue data have been generated and presented in the form of a family of curves. Commonly known as S/N (or /N) curves, they plot applied stress (S) vs. cycles to failure (Nf, or just N) or strain () vs. cycles to failure (N). S/N curves are generated when stress is maintained at a constant level irrespective of specimen condition, whereas the /N curves are generated using test machines that control the total strain of the specimen. S (more commonly known as , the stress range) is the applied stress and is the strain experienced due to the applied stress. Similarly, is simply written in place of the strain range, . Generally, stress is controlled for high cycle-low temperature fatigue, and strain is controlled for low-cycle/high temperature fatigue. These strain curves are asymptotic with fewer cycles to failure at higher stresses or strains. The curves are helpful to designers who may plan components within the threshold where failure is not likely to occur. For engineer-

ing steels at low applied stresses (see Figure 1), this threshold is termed the fatigue limit. For many other non-ferrous materials, failure would occur even at low stresses indicating that the phenomenon of fatigue is material-specific.

Average Stress Amplitude, S

A. Ferrous, Titanium Alloys B. Non-Ferrous Alloys

Design criteria are specified below the curve (threshold)

103

104 105 106 107 108 109 Cycles to Failure, N (Logarithmic Scale)

1010

Figure 1: Typical S/N Curves for Metal Alloys When fatigue stresses operate in the plastic region of a material, plastic deformation induces cyclic hardening or softening of the material. When the material or testing is strain-controlled, cyclic hardening requires increased stress for generating the same amount of strain every cycle. In stress-controlled testing, there is a reduction in the level of strain. A parallel effect is observed for materials that exhibit cyclic softening. These phenomena indicate that as hardening and softening progress in materials, their deformation characteristics vary per their inherent natures and that damage accumulation and spent life are not linear relationships. Fatigue Failure Fatigue failure occurs in three stages. In general, fatigue cracks initiate at the surface due to dislocations which accumulate near stress concentrations. As the material hardens around the stress concentration, cracks nucleate through the brittle region. Removal of dislocations may cause relief of lattice strains or strain softening. The cracks nucleate through the material along the planes of high shear stress. This is called Stage I crack growth. Stage II crack growth involves most of the propagation of crack through the bulk material in a direction perpendicular to the tensile stress. According to some models, dislocation cross slip is considered to be important. In this stage, the extrusions formed on the surface are emitted thus developing into fissures called intrusions. In Stage III, the final stage, the crack has grown to such a large size that the remaining ligament is unable to support it and as a result the final failure takes place in a

AMPTIAC is a DOD Information Analysis Center Administered by the Defense Information Systems Agency, Defense Technical Information Center and Operated by IIT Research Institute

MaterialS E A
conventional manner through the nucleation, growth, and coalescence of micro-voids. The crack propagates in a zigzag transgranular path, perpendicular to the maximum tensile stress. Approximately 99% of the fatigue life is spent on developing fissures into cracks and complete fracture. The factors influencing the rate of crack propagation affect fatigue the most. Figure 2 illustrates the three stages of fatigue crack propagation[2]. STAGE I STAGE II STAGE III
Failure

E
plished by using the right materials (with high fatigue strength or surface residual compressive stress) or by prudent design practices (eliminating corners, notches, and other irregularities which promote cracks). In metallic materials, there is a direct correlation between the tensile and fatigue strengths. Materials with higher tensile strength also possess higher fatigue strength. Besides strength, there are other microstructural factors that affect fatigue properties. In alloys, where a second phase is used for strengthening, spherical dispersion of the second phase provides superior fatigue properties at the same strength level. Alloys of identical chemical composition can have significantly different fatigue limits depending on the sizes of their grains. The fatigue limit is proportional to the square root of grain diameter, which is also the relationship that exists between the yield strength and grain size. In the case of pure metals, fatigue strength is low. To improve fatigue and tensile strength of metals they are alloyed with other elements. Interstitials such as C and N enhance fatigue strength even further by increasing strainaging capability. High Temperature Fatigue Generally, materials lose ductility as temperature decreases and as a consequence, yield and fatigue strength increase. Fatigue strength, however, increases faster than tensile strength. Conversely, an increase in temperature generally reduces the fatigue strength. An example of an alloy that is unaffected by temperature increases is Ni-Cr-Mo steel SAE 4340. The fatigue strength of SAE 4340 remains stable up to approximately 470C. Further increases in temperature cause decreases in fatigue strength. Such behavior points out that accelerated testing for fatigue resistance does not always provide accurate results if temperature is the accelerated factor.[3] Low-cycle fatigue is characterized by the presence of rate (time) dependent effects in addition to the plastic behavior in the high-cycle fatigue range. Tests for low-cycle fatigue resistance are carried out in a strain-controlled environment that simulates the effect of the elastic stress fields surrounded by inelastically deformed regions. At sufficiently high temperatures, creep sets in. Creep failure is different from fatigue, and at intermediate temperatures, there is a creep-fatigue interaction. Creep failure is governed by its own mechanism. The accumulation of slip-lines that initiate Stage I cracking become blunt at higher temperatures, and are thus unable to initiate cracks. Ferrous materials exhibiting fatigue limits at low temperature lose fatigue strength.[4] Fatigue tests at high temperatures are also dependent on cycle frequency, so both time and number of cycles to failure need to be reported for those tests. However, as frequency increases at a constant elevated temperature, the failure becomes independent of frequency. In general, higher creep strength also means higher fatigue strength at high temperatures. Nevertheless, if the room temperature creep strength is enhanced due to finer grain size, the material loses its high temperature creep capability. As temperature increases, the creep resistance of the fine grain structure degrades.[5] Power generation equipment, jet engines, and high temperature components undergo initial changes with temperature increases, from idle to oper-

Shear Plane

Formation of Perpendicular Cracks

Void Growth

99% of Total Service Life

Figure 2: Stages of Crack Propagation across a Specimen[1]. The relative proportion of the three cycles to failure that may be attributed to each crack propagation stage is mostly dependent on test conditions and material composition. However, it is well established that crack initiation generally begins during the first decile of a components life. The majority of crack life is occupied in Stage II, its propagation phase. Stage III, the final fracture time, is relatively short. Ultimate failure (fracture) is caused by the simultaneous action of cyclic stress, tensile stress and plastic strain. If any of the three factors is absent, the crack does not propagate. When design is based on conventional fatigue limit values, care should be taken to consider other geometric factors such as the stress concentration, size, and surface effects. Presence of notches (even a semi-circular notch) or a change in geometry of the component can act as stress raisers, elevating the stress level experienced at particular locations. Larger-sized samples experience a reduction of fatigue strength for probabilistic reasons: a larger sample simply has more flaws to initiate cracks. Surface effects are significant in fatigue since fatigue failure starts from the surface. They are manifested in three different manners - surface roughness, residual stresses and fatigue resistance of the surface region. Variation in surface roughness leads to the variation in stress concentration that in turn affects crack initiation. Residual stresses either due to machining, cold working or heat treatment can add to the applied stress to increase the level of the stress responsible for crack initiation. The fatigue limit of the surface region can change due to loss or enrichment of certain alloying elements or impurities. If this change is conducive to the nucleation of fatigue cracks, fatigue life of the component is reduced. Decarburization, corrosion, oxidation and hydrogen adsorption, are examples of such surface effects. Microstructural Effects A common goal when improving materials is to develop ones with high fatigue strength. For component design, high fatigue tolerance is accom-

A D VA N C E D M AT E R I A L S

AND

PROCESSES TECHNOLOGY

A MPTIAC

tn = Hold Time Stress Range max Stress Range t tension c min c Output Time Input

it is important to test a large random sample. Also, service environments such as lubricating oil, operating temperatures, and a corrosive atmosphere add to the variability of property data, as the they typically are accounted for in laboratory testing. Finally, it is important to define the fatigue failure according to the test program. Definition of failure should reflect failure in service. In some cases, failure is the appearance of the visible crack. In other cases, fatigue failure is the complete fracture of the component. There are several ASTM standards available for fatigue testing, primarily for metallic materials. These standards are shown in Table 1. Table 1: ASTM Standards Available for Fatigue Testing

Ratcheting Test Standard Title Presentation of Constant Amplitude Fatigue Test Results for Metallic Materials Statistical Analysis of Linear or Linearized Stress-Life (S/N) and Strain-Life (/N) Fatigue Data Strain-Controlled Fatigue Testing Measurement of Fatigue Crack-Growth Rate Conducting Force-Controlled Constant Amplitude Axial Fatigue Tests of Metallic Materials A Guide for Fatigue Testing and The Statistical Analysis of Fatigue Data Handbook of Fatigue Testing c - Creep Strain p - Plastic Strain

Stress,

p t

ASTM E468 ASTM E739

Strain, Figure 3: Creep Strain Accumulation Cycle[4] ating conditions up to a steady-state situation. As power output and temperature become constant, creep becomes the dominant damaging process. To simulate periods of nearly constant power output, dwell time at constant strain is introduced. At constant strain, relaxation takes place and stress decreases. However, the stress reduction varies from cycle to cycle and steady-state stress may not be achieved. Hysteresis curves are schematic representations of imposed strain conditions, the resultant stress responses, and crossplots of input and output. Figure 3 indicates that creep strain accumulating in each cycle may lead to failure due to excessive elongation.[4] Under cyclical loading-unloading, the real-situation unloading path does not follow the expected path. The difference between the real and the expected strain, c, is the creep or ratcheting that adds up with each cycle, ultimately causing failure. When determining the projected life (cycles to failure) of materials undergoing cyclic loading, it is important to take into account loading rate and hold time. Cycles to failure are determined by the percentage of load drop. Laboratory Testing Laboratory fatigue tests consist of repeated loading of mechanical test specimens in a load testing machine. However, the results obtained in the laboratory are only a first approximation, and may not necessarily represent field performance. Ultimately, fatigue tests must be carried out on actual parts or assemblies to obtain definitive answers. The main concern with laboratory testing is relating the results to anticipated field service. If the tested parts are similar to previous production items, the field experience helps in designing a sound laboratory test (e.g., complete machines or subassemblies such as transmissions, engines, and hydraulic systems). However, when a new part is designed the type of applied load must be carefully planned. For example, a shaft designed for torsion, but submitted to excessive bending loads in service, may fail. A gear tooth with a high bending strength may fail in service due to a shaft deflection causing excessive local bearing stresses and pitting failure[6]. Ultimately, parts must be tested with the types of loads that are applied in service. Factors such as material lots, tolerances, and surface finish may cause variation in component properties, so

ASTM E606 ASTM E647 ASTM E466 ASTM STP 91 ASTM STP 566

Fatigue Data Fatigue data provide information regarding causes and effects of fatigue phenomena.[7] Their main objective is to establish a relationship between the applied load, stress, or strain value and the fatigue life of the tested component. Table 2 shows the causes and effects of fatigue. Causes are cyclical stress (stress spectrum and stress distribution) and stressed bodies (materials and structures). Effects include atomic, microscopic, and macroscopic phenomena. Fatigue is a property of crystalline materials that results from the interaction of dislocations activated by cyclical stress. If fatigue data are not available, the S-N curves can predict fatigue properties of materials from static loading data. S-N curves can also convert fatigue Table 2: Fatigue Causes and Effects

Fatigue Causes (Environment)


Cyclic Stress Stressed Bodies
Materials Structures

Stress Spectrum
Stress Distribution

Fatigue Effects
Atomic
Dislocation movements Dislocation multiplication Defect interactions Cross slip

Microscopic
Slip formation Slip saturation Structure deterioration Extrusion & intrusion Energy changes Crack nucleation and growth crystallography

Macroscopic
Crack propagation - Stable stage - Unstable stages - Critical length - Final fracture

for a known condition to the fatigue for an unknown one (e.g., predicting the fatigue life of a large part from the data obtained for a small part). Accelerated Testing Accelerated tests are performed by intensifying, in controlled conditions, one of the environmental parameters which cause actual aging or degradation (e.g. heat, humidity, stress, strain, etc.) in the system being tested. The results of accelerated test are only valid when the failure mechanism is identical to the one that occurs in the normal service environment. While intensifying a parameter may accelerate failure, the results have no applicability to actual service if they dont represent realistic conditions. In the case of fatigue, the test parameters are stress or strain, frequency, and temperature. The environment also has a significant effect, but should not be used as a test parameter because of its unknown synergistic effects. Possible applications for each of the four main parameters used in accelerated testing follow: Stress: When stress is considered as an accelerating factor in testing, it is advisable to maintain the stress levels within the elastic range of the material. This is due to the fact that strain hardening varies with strain level. It is even better if a relationship between stress and cycles to failure exists for similar materials (such as a modified alloys) so that only a few data points are required for extrapolation.

10 - 2

Inter - Trans Granular


870C 760C

Strain: For strain-controlled or low cycle fatigue tests, it is possible to increase strain range and accelerate the failure. Again, this is valid as long as the failure mechanism is unchanged. It is worth pointing out that certain predictive models (such as the Coffin-Manson Law) may not be applicable for high strength materials. Also, ceramic materials may not be tested by this method, as they are generally not strain-controlled. Frequency: Use of increased cyclic loading frequency is perhaps the most suitable method for accelerated testing. This is, however, applicable only to metallic materials since crack-growth in ceramics is frequency-dependent (e.g., silicon nitride[8]). In the case of metallic materials, frequency independence is confined to elastic stress-controlled, high cycle fatigue only. Figure 4 illustrates the effect of frequency on a superalloy at high temperature. It also shows how the failure mechanism changes (from transgranular to intergranular) due to the introduction of creep in fatigue testing for a cobalt-based superalloy. Temperature: Temperature significantly affects all phases of fatigue. However, temperature increases are not recommended for accelerated testing. Temperature increases from a low temperature state reduce hardening at locations in the material where fatigue cracks originate, thus increasing the fatigue strength and giving erroneous life estimates. It can also introduce new degradation mechanisms such as oxidation and corrosion which reduce overall fatigue life. Temperature increase from a high temperature state in low-cycle fatigue activates the creep mechanism. Strain aging or solid-state precipitation are also possible. References [1] Creep And Fatigue In High Temperature Alloys; Bressers, J. Ed., Applied Sciences Publishers Ltd., London, 1981. [2] Dieter, G.E.; Mechanical Metallurgy, McGraw-Hill Series in Materials Science and Engineering, 3rd Ed., 1986. [3] Forrest, P.G., Fatigue of Metals, Pergamon Press, 1962 [4] Krempl, E., Wundt, B.M., Hold-Time Effects in High-Temperature Low-Cycle Fatigue: A Literature Survey and Interpretive Report, STP 489, ASTM, Sept 1971, p. 4 [5] Fatigue Design Handbook, A Guide For Product Design and Development Engineers, ASE, 1968, pp. 39-56 [6] Fatigue Design Handbook, ASE, vol. 4, 1968, pp. 106-111 [7] Yen, C.S., Interpretation of Fatigue Data, Metal Fatigue, John Wiley & Sons, Inc., 1969, pp. 107-138 [8] Cranmer, D.C., Richerson, D.W., Mechanical Testing Methodology for Ceramic Design and Reliability, Eds., Marcel Dekker, Inc, 1998, p. 277 [9] ibid, p. 95

Crack Growth Rate, mm/Cycle

10 - 3

[( ) ]
dQ dt creep

600C

10 - 4

Room Temp. dK = 55 MPa m Q = Crack Length

0.001

0.01

0.1

10

Frequency, (Hz)
Figure 4: Frequency Dependence of Fatigue Crack Growth Rate for a Cobalt Based Superalloy[9].

A D VA N C E D M AT E R I A L S
EMAIL:

AND

M A N U FA C T U R I N G T E C H N O L O G Y

AMPTIAC
PHONE: 315.339.7117 FA X : 3 1 5 . 3 3 9 . 7 1 0 7

http

a m p t i a c @ i i t r i . o rg : / / a m p t i a c . i i t r i . o rg

AMPTIAC is a DOD Information Analysis Center Administered by the Defense Information Systems Agency, Defense Technical Information Center and Operated by IIT Research Institute

Anda mungkin juga menyukai