Anda di halaman 1dari 21

Contents

Abstract ................................................................................................................. 1 INTRODUCTION ................................................................................................ 2 FUNDAMENTALS OF FRACTURE .................................................................. 2 What is Fatigue?.................................................................................................... 2 CYCLIC STRESSES ............................................................................................ 4 THE SN CURVE ................................................................................................ 6 CRACK INITIATION AND PROPAGATION ................................................... 8 FACTORS THAT AFFECT FATIGUE LIFE ................................................... 10 Mean Stress ..................................................................................................... 10 Surface Effects ................................................................................................ 10 Design Factors................................................................................................. 11 Surface Treatments ......................................................................................... 11 ENVIRONMENTAL EFFECTS ........................................................................ 12 Case studies ......................................................................................................... 13 FATIGUE FAILURE OF HOLD-DOWN BOLTS FOR AHYDRAULIC CYLINDER GLAND ..................................................................................... 13 References ........................................................................................................... 19

Figure 1 Variation of stress with time that accounts for fatigue failures. (a) Reversed stress cycle, in which the stress alternates from a maximum tensile stress (+) to a maximum compressive stress of equal magnitude. (b) Repeated stress cycle, in which maximum and minimum stresses are asymmetrical relative to the zero-stress level; mean stress range of stress and stress amplitude are indicated. ........ 4 Figure 2THE SEQUENCE OF PROCESSES DURING FATIGUE OF METALLIC MATERIALS. ...................................... 5 Figure 3 Schematic diagram of fatigue-testing apparatus for making rotating bending tests. .......................... 6 Figure 4 Stress amplitude (S) versus logarithm of the number of cycles to fatigue failure (N) ......................... 7 Figure 5 Fracture surface of a rotating steel shaft that experienced fatigue failure. Beachmark ridges are visible in the photograph. ..................................................................................................................... 8 Figure 6 Transmission electron fractograph showing fatigue striations in aluminum. Magnification Unknown .............................................................................................................................................................. 9 Figure 7 Fatigue failure surface. A crack formed at the top edge. The smooth region also near the top corresponds to the area over which the crack propagated slowly. Rapid failure occurred over the area having a dull and fibrous texture (the largest area). Approximately .5x ................................................ 9 Figure 8 surface finish effect .......................................................................................................................... 10 Figure 9 Demonstration of how design can reduce stress amplification. (a) Poor design: sharp corner. (b) Good design: fatigue lifetime improved by incorporating rounded fillet into a rotating shaft at the point where there is a change in diameter. ......................................................................................... 11 Figure 10 Schematic SN fatigue curves for normal and shot-peened steel. ................................................. 12 Figure 11 Schematic of hold-down and bolts (numbered 1-4 ) . .................................................................. 13 Figure 12 Fracture surfaces of bolts. (a) l#; (b) 2#; (c) 3#. Fracture origins are marked with an arrow. ... 14 Figure 13 Fatigue origin zone of 3# bolt ......................................................................................................... 15 Figure 14 Fatigue striations in fatigue propagation zone of 3# bolt ................................................................ 15 Figure 15 Schematic of fracture position of 1# and 2# bolts. ......................................................................... 16 Figure 16 Forces acting on the bolts ............................................................................................................... 16 Figure 17 Stress acting on the fracture surface for I# and 2# bolts. ................................................................ 17 Figure 18 Comparison of fracture surfaces for (a) tensile fracture, and (b) 3# bolt fatigue fracture. .............. 18

II

Abstract
THE discovery of fatigue occurred in the 1800s when several investigators in Europe observed that bridge and railroad components were cracking when subjected to repeated loading. As the century progressed and the use of metals expanded with the increasing use of machines, more and more failures of components subjected to repeated loads were recorded. By the mid1800s A. Wohler had proposed a method by which the failure of components from repeated loads could be mitigated, and in some cases eliminated. This method resulted in the stress-life response diagram approach and the component test model approach to fatigue design. Undoubtedly, earlier failures from repeated loads had resulted in failures of components such as clay pipes, concrete structures, and wood structures, but the requirement for more machines made from metallic components in the late 1800s stimulated the need to develop design procedures that would prevent failures from repeated loads of all types of equipment. This activity was intensive from the mid-1800s and is still underway today. Even though much progress has been made, developing design procedures to prevent failure from the application of repeated loads is still a daunting task. It involves the interplay of several fields of knowledge, namely materials engineering, manufacturing engineering, structural analysis (including loads, stress, strain, and fracture mechanics analysis), nondestructive inspection and evaluation, reliability engineering, testing technology, field repair and maintenance, and holistic design procedures. All of these must be placed in a consistent design activity that may be referred to as a fatigue design policy. Obviously, if other time-related failure modes occur concomitantly with repeated loads and interact synergistically, then the task becomes even more challenging. Inasmuch as humans always desire to use more goods and place more demands on the things we can design and produce, the challenge of fatigue is always going to be with us. Until the early part of the 1900s, not a great deal was known about the physical basis of fatigue. However, with the advent of an increased understanding of materials, which accelerated in the early 1900s, a great deal of knowledge has been developed about repeated load effects on engineering materials. The procedures that have evolved to deal with repeated loads in design can be reduced to four: The stress-life approach The strain-life approach The fatigue-crack propagation approach The component test model approach

INTRODUCTION
The failure of engineering materials is almost always an undesirable event for several reasons; these include human lives that are put in jeopardy, economic losses, and the interference with the availability of products and services. Even though the causes of failure and the behavior of materials may be known, prevention of failures is difficult to guarantee. The usual causes are improper materials selection and processing and inadequate design of the component or its misuse. It is the responsibility of the engineer to anticipate and plan for possible failure and, in the event that failure does occur, to assess its cause and then take appropriate preventive measures against future incidents.

FUNDAMENTALS OF FRACTURE
Simple fracture is the separation of a body into two or more pieces in response to an imposed stress that is static (i.e., constant or slowly changing with time) and at temperatures that are low relative to the melting temperature of the material. The applied stress may be tensile, compressive, shear, or torsional; the present discussion will be confined to fractures that result from uniaxial tensile loads. For engineering materials, two fracture modes are possible: ductile and brittle. Classification is based on the ability of a material to experience plastic deformation. Ductile materials typically exhibit substantial plastic deformation with high energy absorption before fracture. On the other hand, there is normally little or no plastic deformation with low energy absorption accompanying a brittle fracture. Any fracture process involves two steps crack formation and propagation in response to an imposed stress. Fatigue is a form of failure that occurs in structures subjected to dynamic and fluctuating stresses (e.g., bridges, aircraft, and machine components). Under these circumstances it is possible for failure to occur at a stress level considerably lower than the tensile or yield strength for a static load.The term fatigue is used because this type of failure normally occurs after a lengthy period of repeated stress or strain cycling. Fatigue is important inasmuch as it is the single largest cause of failure in metals, estimated to comprise approximately 90% of all metallic failures; polymers and ceramics (except for glasses) are also susceptible to this type of failure.

What is Fatigue?
Fatigue is the progressive, localized, and permanent structural change that occurs in a material subjected to repeated or fluctuating strains at nominal stresses that have maximum values less than (and often much less than) the static yield strength of the material. Fatigue may culminate into cracks and cause fracture after a sufficient number of fluctuations. Fatigue damage is caused by the simultaneous action of cyclic stress, tensile stress, and plastic strain. If any one of these three is not present, a fatigue crack will not initiate and propagate. The plastic strain resulting from cyclic stress initiates the crack; the tensile stress promotes crack growth (propagation). Although compressive stresses will not cause fatigue, compressive 2

loads may result in local tensile stresses. Microscopic plastic strains also can be present at low levels of stress where the strain might otherwise appear to be totally elastic. During fatigue failure in a metal free of crack like flaws, micro cracks form, coalesce, or grow to macro cracks that propagate until the fracture toughness of the material is exceeded and final fracture occurs. Under usual loading conditions, fatigue cracks initiate near or at singularities that lie on or just below the surface, such as scratches, sharp changes in cross section, pits, inclusions, or embrittled grain boundaries. Micro cracks may be initially present due to welding, heat treatment, or mechanical forming. Even in a flaw-free metal with a highly polished surface and no stress concentrators, a fatigue crack may form. If the alternating stress amplitude is high enough, plastic deformation (i.e., long-range dislocation motion) takes place, leading to slip steps on the surface. Continued cycling leads to the initiation of one or more fatigue cracks. Alternately, the dislocations may pile up against an obstacle, such as an inclusion or grain boundary, and form a slip band, a cracked particle, decohesion between particle and matrix, or decohesion along the grain boundary. The initial cracks are very small. Their size is not known well because it is difficult to determine when a slip band or other deformation feature becomes a crack. Certainly, however, cracks as small as a fraction of a micron can be observed using modern metallographic tools such as the scanning electron microscope or scanning tunneling microscope. The micro cracks then grow or link up to form one or more macro cracks, which in turn grow until the fracture toughness is exceeded. The fatigue failure process thus can be divided into five stages: 1. 2. 3. 4. 5. 6. Cyclic plastic deformation prior to fatigue crack initiation Initiation of one or more microcracks Propagation or coalescence of microcracks to form one or more Microcracks Propagation of one or more macrocracks Final failure

These stages in the process of fatigue failure are complicated and are influenced by many factors. This introductory article summarizes some fundamental aspects of the first four stages, prior to the fifth stage of final failure. Further coverage on both crack initiation and crack propagation is given in more detail in subsequent articles in this Volume. In particular, other articles in this section provide detailed information on the fundamental and applied aspects of fatigue crack propagation. Nonetheless, this article briefly considers crack propagation with particular emphasis on the measurement and relation of plastic work with fatigue crack propagation. This article refers primarily to ambient air or an inert environment. The variability in fatigue behavior in similar specimens or parts is well known. This is true for cycles to failure, cycles to crack initiation, micro crack propagation rate, and macro crack propagation rate. A fatigue database should include this variability, as well as a clear description of the specimen or part geometry, surface finish, microstructure, and stress state. The design engineer needs to know about such variability in order to predict probability of failure for a given design. Often reported data do not contain sufficient information about variations in fatigue properties to be useful in a database used for design engineers. Also S-Nf or p-Nf curves of smooth (un notched) specimens are often presented. Macro fatigue crack propagation rates are also presented as smoothed curves based on averaged data points; even when data points are plotted, they represent a trailing average of three to seven points. A fatigue crack does not propagate at a uniformly increasing rate, but frequently slows down or speeds up due to local conditions along the crack front. Real-world spectrum loading further complicates matters. 3

As the resolution of inspection instruments has increased, the portion of the fatigue lifetime ascribed to fatigue crack "initiation" has decreased. For practical purposes, it is useful to define the "initiation stage" as that portion of the fatigue lifetime before a crack is detectable by usual nondestructive evaluation techniques. This is typically to 1 mm. Initiation of such an "engineering" crack represents the preponderance of the fatigue lifetime, except in the regime of very low cycles to failure. After formation of such an engineering crack, the remaining fatigue lifetime is usually relatively short. Thus, the design engineer can design with more precision, utilizing cycles to formation of a small crack combined with Crack propagation to failure, than with cycles to failure alone or with propagation rates of macro fatigue cracks alone. More data that include statistical variations on formation of small cracks in a form useful to the design engineer are needed.

CYCLIC STRESSES
The applied stress may be axial (tension-compression), flexural (bending), or torsional (twisting) in nature. In general, three different fluctuating stresstime modes are possible. One is represented schematically by a regular and sinusoidal time dependence in Figure 1.a, wherein the amplitude is symmetrical about a mean zero stress level, for example, alternating from a maximum tensile stress( max) to a minimum compressive stress ( min ) of equal magnitude; this is referred to as a reversed stress cycle.

Figure 1 Variation of stress with time that accounts for fatigue failures. (a) Reversed stress cycle, in which the stress alternates from a maximum tensile stress (+) to a maximum compressive stress of equal magnitude. (b) Repeated stress cycle, in which maximum and minimum stresses are asymmetrical relative to the zero-stress level; mean stress range of stress and stress amplitude are indicated.

Another type, termed repeated stress cycle, is illustrated in Fig (1.a); the maxima and minima are asymmetrical relative to the zero stress level. Finally, the stress level may vary randomly in amplitude and frequency, also indicated in Fig (1.b); are several parameters used to characterize the fluctuating stress cycle. The stress amplitude alternates about a mean stress m defined as the average of the maximum and minimum stresses in the cycle, or

Furthermore, the range of stress r is just the difference between and max-min namely, Stress amplitude a is just one half of this range of stress, or

Finally, the stress ratio R is just the ratio of minimum and maximum stress amplitudes:

By convention, tensile stresses are positive and compressive stresses are negative.The scheme represents a very simplified description of the real situation, it applies to the behavior of many metallic materials under cyclic loading conditions and allows us to define the term cyclic stress strain behavior (or cyclic deformation behavior). According to Fig. (2), this expression means the area of fatigue that comprises all aspects that deal with the global mechanical and microstructural response of a material to cyclic loading without giving much consideration to damage processes, which are mostly localized. Nevertheless, it should be emphasized that fatigue damage usually evolves from microstructural changes as a consequence of cyclic plastic deformation, and that the study of the cyclic stress-strain response and its correlation with microstructure leads to a better understanding of the early stages of the fatigue failure process.

Figure 2THE SEQUENCE OF PROCESSES DURING FATIGUE OF METALLIC MATERIALS.

THE SN CURVE
As with other mechanical characteristics, the fatigue properties of materials can be determined from laboratory simulation tests. A test apparatus should be designed to duplicate as nearly as possible the service stress conditions (stress level, time frequency, stress pattern, etc.). A schematic diagram of a rotating-bending test apparatus, commonly used for fatigue testing, is shown in Fig (3); the compression and tensile stresses are imposed on the specimen as it is simultaneously bent and rotated. Tests are also frequently conducted using an alternating uniaxial tension-compression stress cycle.

Figure 3 Schematic diagram of fatigue-testing apparatus for making rotating bending tests.

Two distinct types of SN behavior are observed, which are represented schematically in (Figure 3). As these plots indicate, the higher the magnitude of the stress, the smaller the number of cycles the material is capable of sustaining before failure. For some ferrous (iron base) and titanium alloys, the SN curve Fig (4.a) becomes horizontal at higher N values; or there is a limiting stress level, called the fatigue limit (also sometimes the endurance limit), below which fatigue failure will not occur. This fatigue limit represents the largest value of fluctuating stress that will not cause failure for essentially an infinite number of cycles. or many steels, fatigue limits range between 35% and 60% of the tensile strength. Most nonferrous alloys (e.g., aluminum, copper, magnesium) do not have a fatigue limit, in that the SN curve continues its downward trend at increasingly greater N values (Figure 3.b). Thus, fatigue will ultimately occur regardless of the magnitude of the stress. For these materials, the fatigue response is specified as fatigue strength, which is defined as the stress level at which failure will occur for some specified number of cycles (e.g., 107 cycles). The determination of fatigue strength is also demonstrated in Fig (4.b). Another important parameter that characterizes a materials fatigue behavior fatigue life Nf. It is the number of cycles to cause failure at a specified stress level, as taken from the SN plot Fig (4.b).

Figure 4 Stress amplitude (S) versus logarithm of the number of cycles to fatigue failure (N) for (a) a material that displays a fatigue limit, and (b) a material that does not display a fatigue limit.

CRACK INITIATION AND PROPAGATION


The process of fatigue failure is characterized by three distinct steps: 1. Crack Initiation. 2. Crack propagation. 3. Final failure. The region of a fracture surface that formed during the crack propagation step may be characterized by two types of markings termed beachmarks and striations. Both of these features indicate the position of the crack tip at some point in time and appear as concentric ridges that expand away from the crack initiation site(s), frequently in a circular or semicircular pattern. Beachmarks (sometimes also called clamshell marks) are of macroscopic dimensions Fig. (5), and may be observed with the unaided eye. These markings are found for components that experienced interruptions during the crack propagation stage for example, a machine that operated only during normal work-shift hours. Each beachmark band represents a period of time over which crack growth occurred. On the other hand, fatigue striations are microscopic in size and subject to observation with the electron microscope (either TEM or SEM). Fig. (6) is an electron fractograph that shows this feature. Each striation is thought to represent the advance distance of a crack front during a single load cycle. Striation width depends on, and increases with, increasing stress range.

Figure 5 Fracture surface of a rotating steel shaft that experienced fatigue failure. Beachmark ridges are visible in the photograph.

Figure 6 Transmission electron fractograph showing fatigue striations in aluminum. Magnification Unknown

One final comment regarding fatigue failure surfaces: Beachmarks and striations will not appear on that region over which the rapid failure occurs. Rather, the rapid failure may be either ductile or brittle; evidence of plastic deformation will be present for ductile, and absent for brittle, failure. This region of failure may be noted in Fig. (7).

Figure 7 Fatigue failure surface. A crack formed at the top edge. The smooth region also near the top corresponds to the area over which the crack propagated slowly. Rapid failure occurred over the area having a dull and fibrous texture (the largest area). Approximately .5x

FACTORS THAT AFFECT FATIGUE LIFE


The fatigue behavior of engineering materials is highly sensitive to a number of variables. Some of these factors include mean stress level, geometrical design, surface effects, and metallurgical variables, as well as the environment. This section is devoted to a discussion of these factors and, in addition, to measures that may be taken to improve the fatigue resistance of structural components.

Mean Stress
The dependence of fatigue life on stress amplitude is represented on the S N plot. Such data are taken for a constant mean stress m often for the reversed cycles Situation (m=0). Mean stress, however, will also affect fatigue life; this influence may be represented by a series of SN curves at a, each measured different m increasing the mean stress level leads to a decrease in fatigue life.

Surface Effects
For many common loading situations, the maximum stress within a component or structure occurs at its surface. Consequently, most cracks leading to fatigue failure originate at surface positions, specifically at stress amplification sites. Therefore, it has been observed that fatigue life is especially sensitive to the condition and configuration of the component surface. Numerous factors influence fatigue resistance, the proper management of which will lead to an improvement in fatigue life. These include design criteria as well as various surface treatments.
Figure 8 surface finish effect

10

Design Factors
The design of a component can have a significant influence on its fatigue characteristics. Any notch or geometrical discontinuity can act as a stress raiser and fatigue crack initiation site; these design features include grooves, holes, keyways, threads, and so on. The sharper the discontinuity (i.e., the smaller the radius of curvature), the more severe the stress concentration. The probability of fatigue failure may be reduced by avoiding (when possible) these structural irregularities, or by making design modifications whereby sudden contour changes leading to sharp corners are eliminatedfor example, calling for rounded fillets with large radii of curvature at the point where there is a change in diameter for a rotating shaft Fig. (8).

Figure 9 Demonstration of how design can reduce stress amplification. (a) Poor design: sharp corner. (b) Good design: fatigue lifetime improved by incorporating rounded fillet into a rotating shaft at the point where there is a change in diameter.

Surface Treatments
During machining operations, small scratches and grooves are invariably introduced into the workpiece surface by cutting tool action. These surface markings can limit the fatigue life. It has been observed that improving the surface finish by polishing will enhance fatigue life significantly. One of the most effective methods of increasing fatigue performance is by imposing residual compressive stresses within a thin outer surface layer. Thus, a surface tensile stress of external origin will be partially nullified and reduced in magnitude by the residual compressive stress. The net effect is that the likelihood of crack formation and therefore of fatigue failure is reduced. Residual compressive stresses are commonly introduced into ductile metals mechanically by localized plastic deformation within the outer surface region. Commercially, this is often accomplished by a process termed shot peening. Small, hard particles (shot) having diameters within the range of 0.1 to 1.0 mm are projected at high velocities onto the surface to be treated. The resulting deformation induces compressive stresses to a depth of between one-quarter and one-half of the shot diameter. The influence of shot peening on the fatigue behavior of steel is demonstrated schematically in Fig. (9).

11

Figure 10 Schematic SN fatigue curves for normal and shot-peened steel.

ENVIRONMENTAL EFFECTS
Environmental factors may also affect the fatigue behavior of materials. A few brief comments will be given relative to two types of environment-assisted fatigue failure: thermal fatigue and corrosion fatigue. Thermal fatigue is normally induced at elevated temperatures by fluctuating thermal stresses; mechanical stresses from an external source need not be present. The origin of these thermal stresses is the restraint to the dimensional expansion and/or contraction that would normally occur in a structural member with variations in temperature. The magnitude of a thermal stress developed by a temperature Change T is dependent on the coefficient of thermal expansion and the modulus of elasticity E according to

Failure that occurs by the simultaneous action of a cyclic stress and chemical attack is termed corrosion fatigue. Corrosive environments have a deleterious influence and produce shorter fatigue lives. Even the normal ambient atmosphere will affect the fatigue behavior of some materials. Small pits may form as a result of chemical reactions between the environment and material, which serve as points of stress concentration and therefore as crack nucleation sites. In addition, crack propagation rate is enhanced as a result of the corrosive environment. The nature of the stress cycles will influence the fatigue behavior; for example, lowering the load application frequency leads to longer periods during which the opened crack is in contact with the environment and to a reduction in the fatigue life.

12

Case studies
FATIGUE FAILURE OF HOLD-DOWN BOLTS FOR AHYDRAULIC CYLINDER

GLAND
A hydraulic-cylinder gland system used in aircraft failed by leaking in the course of a trial run. The gland was fixed with one hold-down and four hold-down bolts. Three of the four bolts broke in service. The bolts were manufactured by turning and threading from 17-4PH steel. The nominal composition of 17-4PH is OCr-17NiltCu-4Nb and typical mechanical properties are yield strength =1200 MPa, tensile strength b = 1300 MPa after solution heat treatment at 1040"C, then water quenching and tempering for 4 h at 495C. This case describes an analysis of the nature and the causes of fracture as well as preventive measures for avoiding fatigue failure of the hold-down bolts. A schematic drawing of the hold-down and bolts is shown in (Fig. 11). The positions of the holddown bolts are indicated by 1#, 2#, 3# and 4#. Each bolt head was cross drilled with two assembly holes at right angles to one another. The 3# bolt broke away in the middle of the threaded portion. The 1# and 2# bolts broke away in the head between the assembly holes and the shoulder transition radius. General views of the fracture surfaces taken in the scanning electron microscope are shown in (Fig. 12).

Figure 11 Schematic of hold-down and bolts (numbered 1-4 ) .

13

Figure 12 Fracture surfaces of bolts. (a) l#; (b) 2#; (c) 3#. Fracture origins are marked with an arrow.

The fracture surface (Fig. 12(a)) of the 3# bolt was characteristic of a typical fatigue fracture, i.e. there was a crack initiation zone, a fatigue crack propagation zone and a final ductile fracture zone. The fatigue crack initiated at one position in the thread root at a machining mark. The crack propagated towards the far edge of the thread. The origin zone was rough (Fig. 13) and had many radial lines. The propagation zone was smooth and there were distinct fatigue striations (Fig. 14). In comparison with the fatigue surface, the final ductile zone was smaller and was around 20% of the total cross sectional area. According to the abovementioned features, the fracture surface is characteristic of fatigue. The final ductile fracture zone was typically dimpled. No material defects were found in the fatigue origin zone. The fracture surfaces of the 1# and 2# bolts initiated at the edges of the assembly holes, as shown in Fig. 12(b) and Fig. 12(c). A schematic of the fatigue fracture sites is shown in (Fig. 15). The cracks obviously propagated towards the root of the bolt until the remaining cross section became unable to support the load and failed by fast fracture. In comparison with the macrofracture surface of 3#. 14

Figure 13 Fatigue origin zone of 3# bolt

Figure 14 Fatigue striations in fatigue propagation zone of 3# bolt

15

Figure 15 Schematic of fracture position of 1# and 2# bolts.

Stress acting on the bolts: Consider that the platform and the shell of the hydraulic cylinder are rigid because of sufficient thickness. The forces acting on the bolts (Fig. 16). are tensile stresses produced by the changes of oil pressure in the working condition.

Figure 16 Forces acting on the bolts

16

Figure 17 Stress acting on the fracture surface for I# and 2# bolts.

Stress (th) acting on the thread root: First, the effect of stress concentration was not considered. The stress acting on the minimum section of the bolt, i.e. the minor diameter, d1 = 4.134 mm, was the maximum. The oil pressure transmitted by the hold-down platform (d2 = 44.17 mm) was borne by four bolts. The maximum oil pressure Pmax was 3.42 MPa. Take the static estimation as follows:

The fatigue strength of the alloy is much greater than th.max and the bolts should not fail from the thread root if there is a good surface finish, low stress concentration and absence of tensile residual stress. The result of tensile tests for the same group of bolts showed that b was around 1260 MPa. This showed that the maximum stress was about 1/13 of tensile strength of material even in the minimum section. Stress acting between assembly hole and bolt root: It is known from (Fig. 15) that I# and 2# bolts fractured at the location between the assembly holes and the bolt root. Assume that the fracture surfaces were circular cones and were acted on by a uniform tensile stress (Fig. 17).max is comparable to the value obtained above for th.max However, the loadbearing area is reduced by the presence of the assembly holes, and there are appreciable stress concentration factors at the various changes in crosssection. However, a tensile test to fracture of the bolt showed that the threaded portion was still the weakest element with respect to static loading (Fig.18).

17

Figure 18 Comparison of fracture surfaces for (a) tensile fracture, and (b) 3# bolt fatigue fracture.

Remedial measures 1. Increase the distance between the shoulder and the assembly holes (dimension h in Fig. 15). 2. Increase the radius of the filler between the head and shank of the bolt. 3. Use thread rolling instead of threading for manufacturing the thread form. 4. Consider the use of forging for producing the basic bolt shape.

18

References
1) William D. Callister, Jr., Materials Science and Engineering , New York, NY: John Wiley & Sons, Inc., (2007). 2) ASM handbook, Fatigue and Fracture, Vol. 19, (1996). 3) D.r.h.jones, Failure analysis case studies II, UK Elsevier Science Ltd, First edition (2001)

19

Anda mungkin juga menyukai