Sunteți pe pagina 1din 75

PRINCIPLES OF FAILURE ANALYSIS

Types of Failure and Stress

Revised by William T. Becker, Ph.D.

Course 0335 Lesson 2

Copyright 2002 by ASM International All rights reserved

No part of this lesson may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the written permission of the copyright owner. Great care is taken in the compilation and production of this lesson, but it should be made clear that NO WARRANTIES, EXPRESS OR IMPLIED, INCLUDING, WITHOUT LIMITATION, WARRANTIES OF MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE, ARE GIVEN IN CONNECTION WITH THIS PUBLICATION. Although this information is believed to be accurate by ASM, ASM cannot guarantee that favorable results will be obtained from the use of this publication alone. This publication is intended for use by persons having technical skill, at their sole discretion and risk. Since the conditions of product or material use are outside of ASMs control, ASM assumes no liability or obligation in connection with any use of this information. No claim of any kind, whether as to products or information in this publication, and whether or not based on negligence, shall be greater in amount than the purchase price of this product or publication in respect of which damages are claimed. THE REMEDY HEREBY PROVIDED SHALL BE THE EXCLUSIVE AND SOLE REMEDY OF BUYER, AND IN NO EVENT SHALL EITHER PARTY BE LIABLE FOR SPECIAL, INDIRECT OR CONSEQUENTIAL DAMAGES WHETHER OR NOT CAUSED BY OR RESULTING FROM THE NEGLIGENCE OF SUCH PARTY. As with any material, evaluation of the material under enduse conditions prior to specification is essential. Therefore, specific testing under actual conditions is recommended. Nothing contained in this lesson shall be construed as a grant of any right of manufacture, sale, use, or reproduction, in connection with any method, process, apparatus, product, composition, or system, whether or not covered by letters patent, copyright, or trademark, and nothing contained in this lesson shall be construed as a defense against any alleged infringement of letters patent, copyright, or trademark, or as a defense against liability for such infringement. Comments, criticisms, and suggestions are invited, and should be forwarded to ASM International.

ASM International Materials Park, OH 44073-0002 www.asminternational.org

Printed in the United States of America

Acknowledgements
Course Revisers
Roy Baggerly, Ph.D., FASM PACCAR Technical Center William T. Becker, Ph.D. Consultant Daniel J. Benac Bryant-Lee Associates Dennis McGarry FTI/SEA Consulting Ronald J. Parrington IMR Test Labs Incorporated William R. Warke, Ph.D., FASM Retired Research Metallurgist

Technical Advisor
Gordon W. Powell, FASM Ohio State University, Professor Emeritus

Technical Reviewers
Debbie Aliya Susan R. Freeman David N. French Larry D. Hanke William T. Kaarlela Arun Kumar, Ph.D. McIntyre R. Louthan, Jr., Ph.D., FASM Kenneth F. Packer, Ph.D., FASM Robert B. Pond, Jr., Ph.D. James J. Scutti, P.E. Sharam Sheybany, Ph.D. Roch J. Shipley, P.E. Thomas J. Steigauf George J. Theus John A. Wilkinson

Course Procreator
Donald J. Wulpi, FASM Metallurgical Consultant

Project Coordinators
Kathleen S. Dragolich Joanne I. Miller

Principles of Failure Analysis

Types of Failure and Stress


esson 1 provides an introduction to the procedures used in obtaining a root-cause explanation for failure. Many factors must be considered in arriving at that explanation, not all of which are technical. This lesson considers only technical issues. Failure of a part or component can occur in several different ways, only one of which is fracture. Nonfracture failure modes include excessive distortion, wear, and corrosion. Excessive distortion is considered in this lesson. Wear is examined in Lesson 5, and corrosion in Lesson 6. Although it is convenient to separate failure modes in this way, it should not be assumed that they always act independently of each other. Furthermore, the interaction is often synergistic. Wear failures may be accelerated by a corrosive environment (corrosive wear). A corrosive environment may accelerate both single overload failures (stress-corrosion cracking) and cyclic-loading failures (stress-corrosion fatigue). This lesson provides necessary background for the lessons in this Course and considers the following topics:

Upon completion of this lesson you should be able to:


Describe the major types of failure Characterize the various types of distortion and ways to prevent distortion Understand the types of stressaxial, bending, and torsionthat can exist in a manufactured component, how those stresses vary across the cross section of a component, and the deformation that occurs as a result of each type of stress Understand the concept of resolution of stresses into shear and normal components Understand the concept of stress concentration Understand how residual stresses are created, how residual stresses affect behavior (beneficial and harmful), and how to remove residual stresses

Imperfections versus defects The necessity for consistency of conclusions based on fracture surface examinations The types of loading to which a material may be subjected (bending, twisting, pulling, and pushing) Failures due to excessive distortion The sources of stress that may be present in a material Failures at elevated temperature

The first concept to understand is what is meant by the term failure as well as connotations of the term when used in reports and in discussion. Failure does not always involve fracture of a component, although discussions of

Causes of Failure

2 Types of Failure and Stress

failure analysis often emphasize this type of failure. It is better to think of failure as the inability of a machine component to function in an intended manner. A more complete definition might include the following (Ref 1):

The component becomes completely inoperable; fracture need not have occurred. The component operates but does so inefficiently. The component operates but with excessive maintenance. Operation of the part is unreliable. Operation of the component presents a safety hazard.

Failure need not be bad or detrimental. A common type of failure is excessive wear. If the component has served its intended purpose over its intended (hopefully long) design life, failure just means that the component wore out. One does not expect light bulbs to have infinite life, nor does one expect an automobile to function forever. Consumers hope to improve component life by proper and timely maintenance. Failures may occur due to design errors (stress analysis, material selection, geometric sizing), service conditions (loading conditions, environmental conditions), fabrication errors (plastic forming, heat treatment, welding), and assembly errors (galvanic corrosion, fretting fatigue, interference stresses). Local imperfections, which may be geometric features (nicks, gouges) or undesirable microstructures due to improper processing, may be present in the material. These imperfections or deviations from ideality are often found during the course of a failure analysis. Whether they become defects and the cause for failure is one of the essential questions that the failure analyst must answer. A quench crack is an example of an imperfection created during thermal processing when residual stresses cannot be relieved by plastic deformation. If failure of the component resulted in fracture and the crack that led to fracture did not propagate from the quench crack, the quench crack is an imperfection, but not a defect. Imperfections created in one stage of manufacture may or may not be removed in a later stage. Examples include the formation of internal cavities during forging that may close during further deformation, shrinkage porosity in an ingot that may or may not be closed during subsequent working, and the creation of laps during forging. When failure occurs, there are two basic questions that must be answered:

Was the component abused (loaded above its design limit, used for an unintended purpose)? Was the design procedure and manufacturing sequence performed properly?

The purpose of failure analysis is to answer these questions. It must be emphasized that the process of the failure investigation may well, and usually does, reveal some condition in the material that is not ideal, and we

Causes of Failure 3

say that an imperfection is present. The question then is Is the imperfection a defect? Some variability is permitted in all specifications related to producing the final manufactured component, and a factor of safety is included in the stress analysis. It is possible for a processing variable not to be optimum, resulting in, for example, a yield strength on the low side of the permissible range. If this condition is coupled with a service environment variable that happens to be on the high side, say the service temperature, the combination of the two variables can result in failure. Many material specifications include requirements of cleanliness (inclusions) and grain size. Inclusions and grain size in turn affect yield strength, toughness, and ductility. If failure occurs due to mechanical or thermal loading, a successful failure analysis will identify whether failure occurred because the permissible stresses determined in the design process were exceeded, or whether the microstructure and associated mechanical properties assumed in the design process were not present. Permissible stresses could be exceeded because design loads were exceeded, because dimensions of the part were incorrect, or because a local material or geometric imperfection caused the local stress to exceed the failure stress. Some care should be used in the choice of the terms defect and imperfection. If the imperfection is the cause for failure, it can only then be identified as a defect. As noted above, analytical design procedures normally incorporate a factor of safety. The factor of safety is supposed to account for some of the variability in material properties and unknowns in analytical procedures. As a consequence, fairly significant imperfections must usually be present to cause failure in the absence of abuse. Improper microstructures can be obtained because the material was incorrectly processed during fabrication or because the operating environment altered the microstructure. This suggests that a successful failure analysis will (1) identify whether the microstructure of the as-processed component was correct, (2) determine whether the microstructure was altered by the service conditions, and (3) identify whether the design stresses were exceeded. The design process used to determine permissible stresses assumes a specific failure mode. If the component fails in a way not considered during the design process, the implication is that the part was overloaded differently from that assumed in the design and, by implication, the failure occurred at a load lower than the design load. A critical step in failure analysis is to determine the failure mode to be expected from customary usagethat is, the one assumed in the design process. This is done by associating the macroscale and microscale appearance of the fracture surface with a given type of loading. If a particular appearance is not observed, the implication is that the part was loaded in a manner not considered in the design. A few examples are given below. Unfortunately, screwdrivers are sometimes used as chisels or as pry bars. The failure analysis should be able to determine whether the screwdriver was abused by inappropriate use or whether it failed in a way consistent with proper use. Use as a chisel implies axial impact load and

4 Types of Failure and Stress

potential spalling. Use as a screwdriver implies twisting loads. Use as a pry bar implies a bending load. Twisting and bending produce different fracture surfaces. Long-life cyclic load failures occur at stresses lower than that required to cause net-section permanent deformation in a part. If the failed part was designed assuming that failure is by net-section permanent distortion, but there is no macroscopic permanent deformation visible in the part, the part may have failed due to cyclic loading. When a part is cyclically loaded to fracture, a crack initiates and grows in the specimen, reducing the remaining, load-carrying cross section. The part will then fail by overload when the remaining ligament reaches a critical size. There is a characteristically different appearance of the fatigue and overload areas on the fracture surface. Additionally, there are sometimes, but not always, characteristic markings in the fatigue region at both the macroscale and microscale levels of examination. If a part is suspected of failing due to cyclic loading, macroscopic examination may not provide positive evidence for this mode of failure, but the required information may then be provided by microscale examination. Unfortunately, it is also possible that neither macroscopic nor microscopic examination can provide positive evidence of cyclic loading. If a part fails by fracture, it is preferred that the fracture occur in a ductile rather than brittle way because that often precludes a cataclysmic event with no warning that fracture it is imminent. Therefore, it is necessary to be able to identify whether the fracture is ductile or brittle. This information is obtained by examination of both the macroscale and microscale topographical features of the fracture surface. Fractographic information (both macroscale and microscale) must be consistent with the assumed loading conditions and the microstructure of the material. The ideal situation is when both the macroscale and microscale appearances provide positive information regarding the failure process. That is not always the case. Sometimes additional positive information may be obtained by preparing a metallographic specimen normal to the fracture surface. This procedure, for example, is a convenient way to identify intergranular fracture and whether small amounts of plastic deformation accompanied the fracture process. The procedure can also show if inclusions present in the material played an important role in the fracture process. As noted previously, a successful failure analysis of a broken component is obtained when a self-consistent set of conclusions can be drawn from the macroscopic and microscopic appearance of the broken component. This may require all of the following: 1. Recognition of obvious macroscale permanent deformation associated with fracture. Long-life, cyclic-load failures occur at stresses lower than that required to cause net-section permanent deformation in a part. If the part is designed assuming that failure is by netsection permanent distortion, but there is no macroscopic perma-

Causes of Failure 5

nent deformation visible in the part, the part may have failed due to cyclic loading. 2. Observation of the orientation of the fracture surface relative to the component geometry to determine the loading conditions causing fracture and whether the fracture is macroscopically ductile or brittle. 3. Location of the crack initiation site and direction of crack propagation. The location of the crack initiation site(s) can be predicted based on the assumed nominal loading conditions. If the crack initiation site is in a different location, it implies either altered loading conditions, a geometric defect in the partperhaps microscopic in size, or a microstructure that may be generally or locally defective. 4. Identification of the microscopic mechanisms causing fracture. This may be done by examination of the fracture surface at high magnification and/or microstructural examination of a section taken perpendicular to the fracture surface. High-magnification examination of the fracture surface is done using either the scanning electron microscope (SEM) or by using a replica of the fracture surface in either the SEM or the transmission electron microscope (TEM). The TEM is less commonly available than the SEM, and specimen preparation is more tedious. Consequently, TEM examination is less common. However, there are instances in which TEM examination with its higher resolution is required. High-magnification optical examination of the fracture surface is limited to about 1000 diameters and by the depth of field of the microscope. It is therefore of limited use for rough fracture surfaces. 5. Direct or indirect determination of whether the microstructure of the component is consistent with the assumed mechanical properties of the component. In some instances, an inconsistency may be obvious: an annealed microstructure in a presumably cold-worked product or a hot-worked microstructure in a presumably quenchedand-tempered product. In other instances, it may not be so obvious. It is difficult to determine simply by microstructural examination whether a quenched-and-tempered microstructure has been tempered to the correct hardness or whether it has been tempered in an embrittling temperature range. Similar difficulties exist for agehardened microstructures. Sufficient resolution of the microstructure can usually be obtained with an optical microscope. However, in some instances, the microstructure is not resolved at the highest practical optical magnification of about 1000 diameters. It is then necessary to use the SEM to examine the microstructure. Microstructures created by age-hardening treatments and bainitic microstructures created in the heat treatment of steel may require use of the SEM or TEM. Mechanical properties associated with the microstructure in a manufactured component (not necessarily those used in certifying that a component meets some standard) can be directly determined from test specimens taken from the component, and some properties can be inferred from hardness measurements taken on the metallographic specimen used to examine the microstructure. Hardness

6 Types of Failure and Stress

testing can be used to predict both the approximate yield strength and approximate ultimate strength of the component. The latter can be done directly with standard published conversion tables (ASTM E 140). The former usually requires previously determined empirical charts because the variation in yield strength with hardness depends on the microstructure (e.g., the variation is different for say cold working and aging processes or for tempering operations, see Figure 1). Yield strength is more sensitive to alloy composition and microstructure than is tensile strength. However, hardness tests provide little information about ductility or toughness. Microstructural examination of a section normal to the fracture surface will identify whether the microstructure has affected the cracking process. Important microstructural features include whether the cracking is transgranular (TG) or intergranular (IG), whether inclusions or second phases present have influenced the direction of crack growth, grain-boundary composition gradients, and surface composition gradients. The causes for IG fracture are relatively few and often indicate improper processing or a deleterious service condition. A few examples include the fracture of sensitized austenitic stainless steel, temper embrittlement of steels, liquid metal embrittlement, and some combinations of alloy and corrodent that lead to stress-corrosion cracking. Lesson 3 contains a more detailed discussion. The one time when IG fracture is the expected fracture mode is at elevated temperature (temperatures above approximately 40% of the melting point). In some instances, microscale composition gradients can be created in the material by improper thermal processing or inappropriate service conditions, for example, the sensitization of austenitic stainless steels. If present, these gradients can result in fracture adjacent to the grain boundaries at low stresses. Standard etchants used for normal examination of the microstructure may not reveal the presence of the composition gradient. Specific metallographic techniques have been developed for such cases. See, for example, ASTM Standard G 35, Susceptibility of Stainless Steels and Related

Figure 1. (a) Yield strength versus tensile strength and (b) yield strength versus hardness for hot-rolled and quench-and-tempered steels.

Causes of Failure 7

Nickel Chromium Iron Alloys to Stress Corrosion Cracking, and G 36, Stress Corrosion Cracking Tests in Boiling Magnesium Chloride Solutions. 6. It is sometimes fruitful when fracture surface examination, microstructural examination, and hardness tests do not provide consistent and conclusive evidence as to the cause for failure to attempt to duplicate the failure by testing exemplars. If the macroscale and microscale fracture appearances can be duplicated, the root cause for failure has likely been determined. Clearly, collecting all of the above information can be both expensive and time consuming. It is sometimes tempting to omit some of the above testing and examination. Experience with failures of a single component and its associated manufacturing process may provide sufficient familiarity of common causes for failure so that omitting some of the above tests cannot cause an incorrect conclusion to be drawn. The question that should be asked in such an event is What information could be missed if a given test is not performed? Assume that macroscale and microscale examination and/or hardness testing have shown the presence of material imperfections. Unless the imperfection can be shown to be the cause for failure, the imperfection is not a defect. For example, a change in cross section of a component results in a stress concentration at that location, and this location then often becomes the site of crack initiation if fracture occurs. Is the stress concentrator a defect? That depends, and there are two possible answers to this question. Assume that the drawings for the part indicated a minimum fillet radius at the change in cross section. Assume then that the fillet radius measured on the failed part was smaller than the fillet radius called for on the drawing. Is the manufactured fillet a defect? It is a defect in terms of the drawing specification, and, if this question was being litigated in court, the plaintiffs attorney would likely argue that the material was defective because of improper manufacture. However, it has not yet been shown by stress analysis that the decreased fillet radius caused the allowable stress to be exceeded for the specific material (microstructure) in question. If the stress analysis does then show the allowable stress to be exceeded, the imperfection has become a defect and the cause for failure has been identified. Without the stress analysis, all that is known is that the part failed at a location of stress concentration. The location of failure is not at all surprising, and identification of that location as the cause for failure is incorrect. The part will always fail where the internal local stress created by the applied loads and geometry cause the local strength of the material to be exceeded. This lesson discusses the effect of the application of loads and deformations to bodies. It is more convenient to consider the effect of these loads and deformations in terms of the state of stress and strain (distortion) inside the body. We routinely talk about applying a stress to a body, but that is poor and sloppy terminology. Stresses exist only inside a body and are created by the geometry of the component and the magnitude and direction of the applied loads. If two rods having the same microstructure

Concept of Root Cause

What is Stress and What is Strain?

8 Types of Failure and Stress

are pulled to fracture and one rod has a larger cross-sectional area than the other, a larger force is required to break the rod of larger diameter but the stress to cause fracture is the same in both cases. However, if two rods of the same diameter, having different microstructures, are pulled to fracture, they need not necessarily fracture at the same load. In considering these two cases of fracture, it is more helpful to consider the stress required to cause fracture in which case the geometry (i.e., size) of the component need not be considered if the component does not contain any stress risers. The fracture stress is then an inherent property of the material and can be used for comparison purposes with other materials. Stress may be thought of as the intensity of the internally distributed forces, and the stress created inside the body causes a change in shape. Consider an axially loaded member. If the stress is constant at every point in the body, the stress is the load intensity per unit area and is given by the load applied to the body divided by the cross-sectional area. Because the stress in such a case is constant across the cross section, the nominal and local stresses are identical. Changes in cross section (say in a threaded fastener or at a keyway) cause the stress to vary across the cross section, so that we must be clear when talking about the state of stress in a body; that is, whether we mean the nominal stress on the body or whether we mean the local stress at the root of the thread or keyway. The distinction is extremely important because failure is caused by the local stress. Just as stress is related to load via the geometry of the component, strain is related to changes in the dimensions of the body. The strain is defined as a change in length divided by a length, or a change in area divided by an area. Local changes in stress result in local changes in strain and vice versa, so again it is important to know the length or area that is used when the magnitude of a strain is quoted. To minimize possible confusion, Lessons 2 to 4 adopt a labeling procedure that uses S for nominal values of stress, for local stresses, e for nominal strain, and for local strain.

Elastic and Plastic Deformation

A structural member or machine component may be subject to pulling, pushing, twisting, bending, and combinations of these types of loading. Under load, the member or component deforms. For small loads, the deformation is recovered when the load is removed (the part returns to its original dimensions), and the deformation is said to be elastic; that is, the yield strength (YS) has not been exceeded (Figure 2). In the elastic region, the extension is proportional to the force and the strain is proportional to the stress. The proportionality constant for the latter is the elastic modulus (E), sometimes identified as Youngs Modulus or the modulus of elasticity. At higher loads, the yield strength is exceeded and the deformation is not completely recovered upon removal of the load. Above the yield stress, deformation is both plastic and elastic under load and the elastic deformation is recovered when the load is removed. Permanent deformation is described as plastic deformation and initiates when the yield strength is exceeded. At still higher loads (starting at maximum load), deformation usually becomes localized in the member and fracture usually occurs soon after localized deformation begins (there are examples of

Causes of Failure 9

765 BBIAJ ;5
2 2K

A? E C IJ=HJI

5 FA
E = IJHAII H =@

2
E = IJH=E = @ ANJA IE

) +

* ,

Figure 2. Stress-strain curve for a tensile specimen. If a stress S1 (load P1 divided by the initial area Ao) does not exceed the yield strength (YS), there is no permanent deformation when the load is removed. For the stress S2 (P2 divided by Ao), the YS (approximate end of the linear region), there is permanent deformation under load, whose magnitude is visible as (A) when the load is removed. The elastic strain (B) is recovered when the load is removed. If the material is loaded above the maximum the strain (change in length divided by original length) becomes localized in the specimen, leading to the formation of a neck. Fracture then occurs in the necked region. When the stress S2 exists in the specimen, the elastic strain energy available to drive a crack is the triangular region S2 high and (C A) wide. UTS, ultimate tensile strength; E, elastic modulus. torsional loading in which there is no visible localized deformation). In many instances, permanent deformation in a material constitutes failure. Examples include permanent deformation of a spring, permanent deformation of a crankshaft or connecting rod, and permanent deformation of a tie rod. However, failure may also occur when the loading creates only elastic deformation as is discussed in the section Stable Deformation and Unstable Deformation (Buckling) in this lesson. Because the fracture initiation site and the resulting fracture surface orientation depend on the kind of loading applied (axial, bending, twisting), it is important to understand the state of stress and strain that is created by these types of loading. The state of stress acting in a body is analyzed in terms of stresses acting parallel to a plane causing sliding (shear stresses) (Figure 3) and stresses acting perpendicular to a plane (normal stresses). Any resultant stress acting on a plane can be decomposed into shear and normal components. Shear stresses cause a change in angle between two lines that were initially perpendicular (Figure 4), but no change in the distance between the planes (not strictly true if the shear strain is large). Axial loading causes no change in angle between originally orthogonal lines, but does

Normal Stress and Shear Stress

10 Types of Failure and Stress

Figure 3. Definitions of normal (axial) stress and shear stress. Normal stresses act perpendicular to a plane and shear stresses act parallel to a plane. If the body shown is cut on an oblique plane, the internal state of stress can be resolved into shear and normal components. The relative magnitude of these components depends on the angle () of the cutting plane to the applied force. Normal stresses may be positive (tensile) or negative (compressive). Positive shear stress causes a counterclockwise rotation, and negative shear stress causes a clockwise rotation (righthand screw convention). change the distance between the lines (planes). If the axial loading is tension (Figure 4), the spacing increases between lines orthogonal to the load and it decreases between lines parallel to the load. Consider the body shown in Figure 3, which has an external tensile load acting on it. If the body is sectioned as indicated in Figure 3, equilibrium requires that there be a stress on the exposed oblique face. This stress can be resolved into components parallel and normal to the exposed plane. The relative magnitudes of the shear and normal components vary with the angle phi (). Analysis shows that the maximum shear stress on the oblique plane occurs when 45 and that the shear stress is then onehalf of the normal stress. If the component is loaded in compression, the same result is obtained. It is assumed when discussing the macroscopic appearance of failed specimens that brittle fracture occurs on planes of maximum normal stress and that ductile failures occur on planes of maximum shear stress.

Normal Strain and Shear Strain

Axial loading causes no change in angle between originally orthogonal lines, but changes the distance between the lines (planes) and therefore creates axial (normal) strain. If the axial loading is tension, the spacing increases between lines orthogonal to the load and it decreases between lines parallel to the load (Figure 4). The opposite is true for compressive loading. Metallic materials deform under conditions of essentially constant volume. Therefore, when a tensile stress is present, the material elongates in the direction of the tensile stress and contracts perpendicular to the tensile stress. The ratio of the distortion in a direction perpendicular to the load to that in the direction of the load is given by Poissons ratio. When material deforms elastically, Poissons ratio is about one-third for most materials, and it increases to about one-half when the material deforms plastically. Many times, fabricated components contain surface

Axial, Shear, and Bending Stresses 11

Figure 4. Definitions of axial or normal strain and shear strain. (a) Tensile axial strain results in an increase in length parallel to the load and contraction perpendicular to the extension. Compressive axial strain results in decrease in distance between lines perpendicular to the load, and an increase in distance between lines parallel to the load. All lines originally at 90 remain at 90 after the deformation. (b) Small shear strain results in a change in angle between two originally orthogonal lines (), but there is no change in distance (h) between parallel planes. markings that result from the fabrication process. Two examples are machining marks and scratches from drawing or extrusion. Careful examination of these marks can determine whether a material was loaded axially, bent, or twisted. Shear strains are created when a body is subjected to a shear stress. Shear deformation is different from axial deformation in that there is a change in angle between originally orthogonal lines, but the distance between parallel lines does not change (at least for small strain) (see Figure 4). Axial loading may be tensile or compressive. Consider the axially loaded member shown in Figure 3. Imagine two sets of scribed lines on the surface of the specimen, one set parallel to the load, and a second set perpendicular to the load. Tensile stresses are present if the specimen elongates in the direction of the load, that is if the distance between parallel planes perpendicular to the load increases. Compressive stresses are present if the specimen becomes shorter, that is if the distance between the scribed lines perpendicular to the load decreases. Note too that lines scribed parallel to the axis of the member do not change orientation relative to the specimen axis during loading (unless the load exceeds the maximum). In the axially loaded member described above, the stress does not vary across the cross section before the maximum load is reached. That indicates that in the absence of an imperfection in the material, brittle fracture would be expected to occur on any plane normal to the load and could initiate at

Axial, Shear, and Bending Stresses


Axial Loading

12 Types of Failure and Stress

any point on that plane. The actual location is often at the surface or along the centerline because these are the two locations where imperfections are most likely to be present in the material. Imperfections may be either macroscale or microscale. They increase the stress locally above the nominal value. Stress concentration is also created at any change in cross section (larger to smaller diameter, fillets, keyways, thread roots, etc.).

Torsional Loading

Twisting and bending loads create a more complex state of stress that varies with position in the body and, in general, that consists of both shear and normal components. Consider the cylindrical specimen shown in Figure 5. If a series of lines parallel and perpendicular to the specimen axis are drawn on the specimen prior to loading, the circumferential lines show no change in orientation when the specimen is loaded, but the axial lines become twisted relative to the specimen axis. For small deformations, the distance between the circumferential lines does not change, indicating that there is no normal strain parallel to the axis. However, because the angle of the longitudinal lines relative to the axis does change, this suggests that there are shear strains and stresses present in the member (compare to Figure 3). There are normal stresses acting in the body, and they can best be visualized by cutting the material on a plane oriented 45 to the axis as shown in Figure 6. If the specimen is now twisted as shown, the cut section opens up, indicating the presence of a tensile stress. However, if it is twisted in the opposite direction, the cut does not open. If the same experiment is repeated except the 45 cut is oriented in the opposite direction, the cut will open if it is twisted in the opposite direction from that shown in Figure 6. This experiment indicates that there are two sets of planes of maximum normal stress in the body

Figure 5. A cylindrical section subjected to torsion loading (T). Circumferential lines (OA) perpendicular to the specimen axis remain perpendicular during loading (OB). Lines parallel to the axis on the surface (CA) become rotated (CB). Compare to Figure 3.

Axial, Shear, and Bending Stresses 13

one set on which a tensile stress acts and another set 90 away on which a compressive stress acts (Figure 7). Analysis also indicates that the magnitude of each of the two normal stresses is the same and that the magnitude of the normal stress is the same as the magnitude of the shear stress. In a cylindrical specimen, there are clearly an infinite number of sets of the plane of maximum normal stress. In the same way, there are an infinite number of the longitudinal planes of maximum shear stress. Any plane parallel to the axis and containing the axis is a plane of maximum shear stress. Analysis of the displacement of points on a transverse plane during loading indicates that the displacement goes to zero along the axis of the cylinder. Because the displacement on a circular arc is given by displacement radius times angle (s r), the strain on the transverse plane varies radially with distance from the axis. It is at a maximum on the surface and zero along the axis. (See Figure 5.) Lines are not normally drawn on manufactured components prior to placing them in service. However, as noted above, many manufacturing operations leave markings on the surface that can be used to identify the type of loading as in the cases described previously. Extrusion processes

Figure 6. Demonstration of stresses created in torsion loading. (a) Shear stress and shear strain on a plane of maximum shear stress. There are two planes of maximum shear stress, one perpendicular to the axis of the cylinder and the other parallel to the axis of the cylinder and containing the axis. (b) A plane of maximum normal (tensile) stress. There are two planes of maximum normal stress, both at 45 to the axis of the cylinder. The normal stress is tensile on one of these planes (as shown), and compressive on the other plane. Source: Ref 2, p 33.

14 Types of Failure and Stress

Figure 7. Planes of maximum shear stress and maximum normal stress in a cylindrical section loaded in torsion. often leave die marks parallel to the extrusion direction. Rolling typically leaves markings parallel to the rolling direction. Machining leaves fine markings perpendicular to the axis of the part. Examination of these lines in the fractured component is helpful in determining the loading conditions that were present at the time of failure.

Bending

Consider the three-point-loaded beam shown in Figure 8. The force P causes the beam to bend downward, creating a tensile stress acting parallel to the axis of the beam on the bottom of the beam because its length is increased and a compressive stress acting parallel to the axis on the top of the beam because its length is decreased (Figure 8a). So long as the beam is deformed elastically, this normal stress varies linearly from a maximum tensile stress on the bottom to a maximum compressive stress on the top of the beam. The location where this stress goes to zero is known as the neutral axis (NA) (Figure 8). The normal stress (tension or compression) may be calculated according to:
S Mc/I (Equation 1)

where M is the bending moment, c is the distance from neutral axis, and I is the moment of inertia. Values of I and the location of the neutral axis for various geometries are tabulated in many handbooks. Equation 1 is sometimes written as:
S M/Z (Equation 2)

Axial, Shear, and Bending Stresses 15

AKJH= =NEI

) N )

*
1B JDA >A= IKFF HJ =J ) EI HA LA@ = @ JDA >A= EI ?KJ =J * JDA IA?JE B JDA >A= ) *  A CJD N KIJ D=LA = IDA=H B H?A 8 =?JE C =J * ) I JDA IDA=H B H?A 8 ?HA=JAI = A J E JDA >A= =J 8 B =C EJK@A 8N
+ FHAIIE 6A IE AKJH= =NEI

8 *

Normal stress (a)

AKJH= =NEI

Shear stress

(b)

(c)

Figure 8. Stresses in a beam loaded in bending. Two stresses are present in the beam. A normal stress that acts parallel to the beam, and a shear stress that acts on planes parallel and perpendicular to the axis of the beam. (a) Origin of shear forces and bending moments in a beam. (b) Origin of longitudinal shear stresses in a beam. (c) Longitudinal shear stresses present in the beam shown in (a). where Z is the section modulus, I/c. Note that the maximum stress is located where the bending moment is a maximum and at a maximum distance from the neutral axis. For a beam symmetrical about the half height (rectangular, circular, I-beam with equal width flanges, etc.), the neutral axis is at half height. There are also shear stresses acting in the beam. They can be visualized by flexing a deck of cards as in Figure 8(b). As the cards are flexed, they slide over each other (Figure 8b, middle sketch). If this sliding is prevented, a shear stress must be present (Figure 8b, bottom sketch). It can be shown that the shear stress reaches a maximum along the neutral axis and goes to zero at the outside surfaces of the beam (Figure 8c). The variation of the shear stress, sometimes called the longitudinal shear, is shown in Figure 8(c). This shear stress is usually not very large and is often neglected

16 Types of Failure and Stress

in a stress analysis. For a rectangular cross section, the maximum shear stress is 4/3 the average shear stress (shear force divided by the cross-sectional area), 3/2 the average shear stress for a circular cross section, and twice the average shear stress for a hollow cylindrical section. However, if the beam is also subjected to twisting forces about the longitudinal axis, the shear stress created by bending adds to the shear stress created by twisting, and, therefore, it should be considered, say for example, in a section of piping subjected to both bending and torsion. As in torsional loading, then, the stress state in the beam varies from point to point. However, for the beam, the longitudinal tensile stress is a maximum where the shear stress is zero, and the shear stress is a maximum where the normal stress is zero. In the case of the cylinder loaded in torsion, the shear stress and normal stress are of equal magnitude at any radial distance from the axis of the cylinder. In most, but not all beams, the magnitude of the shear stress is sufficiently small that it does not cause failure of the beam. A laminated structure consisting of two different types of materials, one of the materials being significantly weaker than the other, could fail due to the shear stress created by flexing. This behavior is not common in metallic materials, but could occur in sweat-soldered joints if the contact area is not large enough. It is a common failure mechanism in nonmetallic materials bonded together. Cyanoacrylate cements are not especially strong in shear, and assemblies bonded by this adhesive do fail in shear when subjected to a bending load. Composite beams created by welding or riveting plates together can also fail in this way if they are not properly designed.

Crack Initiation Sites

Failure analysis requires identification of the crack initiation site. There are multiple ways by which the initiation site can be determined, including use of characteristic markings on the fracture surface, as discussed in Lesson 3. Unfortunately, the presence of these markings depends to some extent on the strength level of the material and also on whether the material is in the cast or wrought condition. Cast materials tend to show less distinctive features on the fracture surface, and it is therefore sometimes necessary to resort to the following method. One expects the crack to initiate on the side of the beam loaded in tension. The tensile side of the beam must contract perpendicular to the internal tensile stress, and the beam must expand on the compression side because of the Poissons ratio, as discussed previously. Therefore, unless the fracture is ideally brittle, the tensile side of the beam can be determined by careful measurements of the transverse section. The thin part of the beam is the side that was loaded in tension. Secondly, failed parts suspected of failing in bending can be placed on a flat surface and brought together (making sure the fracture surfaces do not contact or are rubbed together). A change in angle due to plastic bending is then usually obvious. When fracture in bending is brittle, as say for gray iron castings, careful examination of the edge of the component will also sometimes determine the tension side of the beam. The tension-side edge is rougher than the compression-side edge (Ref 3).

Stress Concentration 17

Nominal states of stress are modified by local changes in geometry, for example at a fillet connecting two cylindrical sections of different diameter (Figure 9), at keyways, and by many other geometries. The local stress at a fillet is greater than the nominal stress (given by the load divided by the cross-sectional area). The ratio of the actual local stress, , to the nominal stress, S, is known as the stress concentration factor, kt, that is:
kt(S) (Equation 3)

Stress Concentration

For fillets, stress concentration factors increase as the difference between the two cross sections increases and as the fillet radius decreases (Figure 9). Stress concentration exists for all types of loadingaxial, torsion, and bending. Most mechanical design books provide some examples of stress concentration factors for various loading geometries. The most complete reference is that due originally to Petersen and subsequently augmented by Pilkey (Ref 4), who has continued Petersens work.

Figure 9. Stress concentration in a section containing a fillet and subjected to axial loading. Source: Ref 4, p 156.

18 Types of Failure and Stress

There is nothing wrong with the presence of stress concentrations in manufactured components caused by deliberate manufactured changes in cross section. Stress concentration factors are an accepted part of engineering design. The issue involving stress concentration is to make sure that the local stress does not become too large. Examination of standard references for stress concentration factors and consideration of accepted design practice reveal that stress concentration factors seldom exceed a magnitude of three in engineering design as, for example, in the case of a round hole in a wide plate. The presence of a sharp, cracklike defect in a member is another matter. Although strictly based on the geometry of an elliptically shaped defect, a reasonable approximation of stress concentration for a long, sharp, cracklike imperfection is given by:
kt smax 12 Snom c r

(Equation 4)

where c is the crack length and is the radius of curvature at the crack tip. Now, the danger inherent in the presence of a quench crack can be seen. The radius of curvature of the crack tip is of the order of the interatomic distance, for example, 2 (2 1010 m, or 2 107 mm). For a quench crack with a length of 0.5 in. (12.7 mm), the resulting stress concentration factor is:
kt 1 2 12.7 mm 15,900 2 10 mm
7

(Equation 5)

No material can withstand that magnitude of stress concentration. Yielding must occur if the material has any inherent ductility, and brittle fracture must occur in an inherently brittle material.

Hertz Contact Stresses

If two spherical balls are placed in contact under load, deformation occurs in both balls so that the initial point of contact becomes a circular area. Similarly, if two cylindrical sections are placed in contact under load, the initial line contact becomes a thin area. A similar situation develops when curved sections are in contact with flat surfaces. Stresses for these geometries can become very large due to the small contact area. Additionally, these stresses are triaxial. Such loading conditions are created in machine components in many instancesthe contact of gear teeth, roller bearings, and cams. A detailed discussion of the state of stress is beyond the scope of this lesson. The equations that describe the magnitude of the triaxial state of stress are messy and depend on the contact force, the difference between the radii of the two cylinders, the tensile modulus, and the Poissons ratio of each cylinder. However, the equations indicate that all of the normal stresses acting on the element are compressive and that the location of maximum compressive stress is at the surface of contact. The maximum shear stress occurs at a point below the surface (Figure 10). Spalling or crushing can occur near the surface due to the compressive stresses, and it is commonly accepted that fatigue failures can be initiated at the location of the maximum shear stress. For the case of two steel cylinders with diameters of 1/8 in. and 1 in. (3.2 and 25 mm), the maximum shear stress occurs at a point 2 to 4 mils (50 to 100 m) below the surface. This distance is not too different from the case depth of a typical nitrided part,

Service at Elevated Temperature 19

Figure 10. Contact stresses between two cylinders. The horizontal axis is plotted in terms of the maximum pressure (p0) at the centerline of contact The vertical axis is plotted in terms of a depth below the surface. See text for discussion. Source: Ref 5, p 324 but it is considerably less than the case depth of a carburized or inductionhardened part. The role of shear stresses in initiating fatigue failures can be determined using Figure 11. The shear stress acting underneath the surface is completely reversed as a given site moves through the arc of contact. There is some tendency for the driven cylinder or gear to slide over the undriven cylinder or gear, the sliding action creating a friction force. This friction force in turn creates normal and shear stresses that are superimposed on the contact stresses shown in Figure 12. These latter stresses are maximum at the surface of the two cylinders. The normal stresses created by the friction also change sign as a point moves through the contact area, thereby subjecting the material to another stress cycle. Finally, any friction caused by sliding at the arc of contact causes local heating. When materials are subjected to elevated-temperature service, several conditions degrading to the behavior of the material can come into play. When a steel part is loaded in tension at room temperature, we expect it to elongate in the direction of the load and for the elongation to increase if the load increases. Further increases in length occur only for additional increases in load. If the load is held at a constant value, there is essentially no change in length. However, if the same experiment is conducted at

Service at Elevated Temperature

20 Types of Failure and Stress

Figure 11. State of stress along the line of contact for two gears. Source: Ref 5, p 326

room temperature with a low-melting material such as lead, we would observe continued change in length with no change in load. That is, deformation in the lead specimen is time dependent as well as load dependent. When materials deform at constant load, they are said to creep (Figure 13). Alternatively, if a member is stretched to some length and then the load required to maintain that length is measured, one finds that the load decreases with time at elevated temperature. Time-dependent load change at constant deformation is termed stress relaxation. Stress relaxation is an important failure mechanism in threaded fasteners. As the component creeps and the stress relaxes, the clamping load is lost on the fastener, so that relative movement between two parts can occur. Loss of clamping load is a common cause for fatigue failures in bolted assemblies. If the clamping force is restored during maintenance, the material is

Service at Elevated Temperature 21

Figure 12. Stresses created by friction at the line of contact between two gears. Source: Ref 5, p 327

Figure 13. Creep curve. At elevated temperature, the time-dependent deformation of a material must be taken into account since deformation will occur at constant load. The curve shown is typical of a material tested under constant load at a homologous temperature of about 0.4 or more. strained. Multiple retightening of the fastener to restore the clamping force under creep conditions uses all of the available strain in the material, and fracture occurs. Several other changes can occur in a material at elevated temperature. The stress to cause yield and the stress to cause fracture decrease as the temperature increases. Strain at fracture goes through a minimum at elevated temperature (with an associated change from TG to IG fracture). The elastic modulus of the material decreases. Surface oxidation may become significant and can result in scaling. In addition, oxide particles may precipitate at the grain boundaries. The surface composition can

22 Types of Failure and Stress

change (for steels, this can be decarburization in an oxidizing atmosphere or carburization in a reducing atmosphere), and the nominal microstructure may change. Eutectic and eutectoid microstructural constituents typically spheroidize (e.g., the cementite in pearlite in steels changes from a lamellar to a spheroidal morphology). A cold-worked material placed in service at elevated temperature may recrystallize, resulting in a decrease in yield strength. Additionally, cementite present in a steel can decompose to graphite. The stability of the carbide depends on alloy composition (whether graphitizers or carbide formers are present). Graphitization in plaincarbon steels initiates at about 800 F (425 C) and accelerates rapidly as the temperature is increased. Alloying elements that lead to graphitization include molybdenum, silicon, and aluminum, which would be present because of deoxidation practice. The addition of chromium to steel not only increases carbide stability, but also prevents scaling, so that many steels selected for high-temperature service (which happen to contain some molybdenum for other purposes) contain chromium. Another common microstructural change associated with elevatedtemperature service is the overaging of age-hardening alloys. Elevated temperature also results in increased corrosion rates and the possibility of liquid metal embrittlement. Types of embrittlement are discussed in Lessons 3 and 6. Each of the degrading conditions discussed in this section has unique kinetics and temperatures at which it becomes important. This then means that the term elevated temperature has to be defined. This is done in terms of the homologous temperatureservice temperature divided by the melting point in degrees absolute). Time-dependent behavior starts at very low temperatures. However, from an engineering point of view, creep mechanisms may be considered important when the homologous temperature exceeds approximately 0.4. (Note that overaging in age-hardening alloys may occur at a lower homologous temperature.) In the remaining sections of this lesson, it is understood that the temperature is too low to permit creep deformation unless it is specifically considered in the discussion.

Macroscale versus Microscale Behavior

The terms ductile and brittle carry different interpretations based on the context in which they are used. For example, both the analytical design procedures and factors of safety that are utilized are different for ductile and brittle materials. Whether a material is ductile might be specified in terms of the tensile elongation or reduction of area obtained from a tensile test: for example, set arbitrarily at 10% tensile elongation. Alternatively, the distinction can be made on the basis of macroscale visible appearance or in terms of microscale deformation mechanisms. There is still one more consideration. When we say that ductile fracture has occurred, do we mean that plastic deformation preceded the fracture process or do we mean that the fracture process itself was ductile?

Macroscale versus Microscale Behavior 23

The macroscale fracture appearance of a component can be brittle, even though significant permanent deformation occurred prior to the fracture process. Although it is considered poor practice because of potential damage to the fracture surface, bringing the two halves of a broken component together to see if the reassembled component shows no distortion can be used to identify ductile and brittle fractures. This is, for example, a sensitive technique to determine if plastic deformation occurred in bending prior to fracture. A related technique is to obtain high-resolution traces of the fracture surfaces and then to superimpose these traces. If they can be brought into superposition, the fracture is brittle. If they cannot, the fracture is ductile. Such a procedure is time consuming and requires training in the use of sophisticated equipment. For many situations, this level of sophistication is not necessary, and more practical and usable definitions of ductile and brittle fractures are desired. There can be some ambiguity in characterizing the fracture process as being ductile or brittle unless the description is done carefully. This is due in part to the observations that (1) the macroscale appearance may be brittle, but the fracture surface is created by ductile, microscale processes not visible to the unaided eye, and (2) the macroscale fracture surface can change from ductile to brittle or from brittle to ductile during crack propagation, and crack propagation can occur by simultaneous ductile and brittle processes on the microscale level. Because of (1), it is necessary to be clear in describing the fracture whether it is the macroscale appearance or the microscale mechanism that is being discussed. At the macroscale, a fracture is described as ductile or brittle based on its appearance. Fractured components showing obvious distortion (including necking) are described as ductile. In the absence of visual distortion, fracture is described as ductile if it occurs on a plane or planes of maximum shear stress and brittle if it occurs on a plane of maximum normal stress. Evidence from microscale examination is used to indicate whether the mechanism causing fracture is ductile or brittle. Such evidence can be obtained by metallographic examination of a section through the fracture surface (bent annealing twins in face-centered cubic materials, strain-induced martensite in an austenitic stainless steel), by direct examination at high magnification of the fracture surface in the SEM, or by examination of a replica of the fracture surface in the TEM or SEM. Macroscale crack initiation and propagation without prior macroscale, visible plastic deformation is described as brittle.* However, brittle fracture may also be initiated after prior, macroscale plastic deformation, that is, brittle fracture in a component subjected to prior cold work. If this is the case, the fracture should not be characterized as brittle without a comment that the component had been cold worked prior to service. Fracture that occurs following macroscale plastic deformation is described as ductile fracture. However, it is also possible for a crack to

*In such an instance, the cleavage crack is formed by dislocation motion and, therefore, from a mechanistic point of view is ductile because dislocation motion implies plastic deformation. However, from an engineering perspective, the fracture is considered brittle.

24 Types of Failure and Stress

propagate partially by a macroscale and microscale brittle mechanism and then to propagate further by a microscale and macroscale ductile process. Additionally, crack propagation may occur simultaneously by brittle and ductile mechanisms (e.g., quasi-cleavage in steels; see Lesson 3). Again, the fracture should not be described simply as being ductile or brittle. Whether the microscale processes causing fracture can be identified without resorting to high-magnification examination depends upon the volume of the material involved in the failure process and, in some cases, depending upon the information obtained from macroscale examination of the fracture surface. If the mechanisms responsible for permanent deformation are restricted to too small a volume, they are not visible macroscopically. This suggests that examination of a fracture should be performed at two magnifications, and that is often necessary. Gross bending distortion of a component is visible with the unaided eye. However, a macroscopically flat fracture surface may be created by microscale processes that are either ductile or brittle. As is developed in this lesson and extensively in Lesson 3, the orientation of the fracture surface with respect to the applied loads provides some clue as to not only the type of loading (e.g., pulling, bending, or twisting), but also whether the failure mechanism is ductile or brittle.

Macroscale Causes for Failure

It was noted in the introduction to this lesson that failure could occur without fracture. We expect fracture to occur when materials are subjected to sufficiently high tensile stresses, as in axial loading, bending, and torsion. However, failure can occur before fracture occurs, for example, by excessive elastic deformation. A first impression might be that fracture cannot occur in compression and that the only possible failure mode is by buckling. However, such is not the case. Compression loading often develops secondary tensile stresses that act perpendicular to the applied compressive stress, and cracking may develop due to these secondary stresses. This situation is discussed in detail in Lesson 3. Brittle materials loaded in compression may fail on a shear plane (as for example, gray cast iron), but brittle fracture can also initiate along the centerline as discussed in Lesson 3. Alternatively, compressive stresses created in axial loading or bending can cause unstable buckling deformation, and buckling may also develop under torsional loading. It is convenient to classify fractures as occurring due to stable or unstable deformation and according to whether fracture did or did not occur. The relationship between loading conditions and fracture surface orientation is discussed in Lesson 3. This lesson considers failure when fracture does not occur and also will consider how residual stresses can be created in a material and their effect on behavior of the material.

Failure in Tensile Loading

The axially loaded tensile member shown in Figure 2 is a convenient way to introduce consideration of design versus material selection/processing issues and to identify several failure modes. If this specimen is pulled to failure and the load is monitored as the specimen extends, the plot of force versus extension is often like that shown in Figure 2.

Failure in Tensile Loading

25

For a sufficiently large load, the tensile member in Figure 2 will fail by fracture. However, it can fail in several other ways at lower loads than that required to cause fracture, and each of these types of failure is controlled by an inherent property of the material. Assume that a load P1 is applied to the member, causing it to extend in the direction of the load. If this load (P1) is not very large, one finds that when it is removed, the specimen returns to its original length. The deformation is described as elastic or recoverable. The load, P, may be converted to a nominal stress by dividing the load by the initial cross-sectional area (Ao), that is:
S P Ao (Equation 6)

Similarly, the observed increase in length may be converted to a strain, e, by dividing the length change by the original length of the specimen. Then, the ratio of the stress to the strain is the modulus of elasticity or Youngs modulus (E):
e L Lo (Equation 7)

S E e

(Equation 8)

The first potential failure mode has now been identified: excessive distortion, that is, excessive elastic strain, for a given design load. The strain could be reduced by decreasing the stress, that is, by either increasing the cross section and/or decreasing the load, or by increasing Youngs modulus. Youngs modulus is a physical property of the material, controlled by interatomic bonding forces, and it is not affected by heat treatment of metallic materials. For common engineering metallic materials, it ranges from about 10 106 psi to about 30 106 psi (69 to 207 GPa), so the ability to decrease the strain by changing materials is limited. The ability to increase the section size to lower the stress may also be limited, say when the design requires many components to be placed within a small volume. Elastic moduli of polymeric materials are considerably smaller than those of metallic materials (a few hundred thousand psi to one or two million psi), so that the stiffness of a component manufactured initially using a polymer could be increased by changing to a metallic material. Elastic strains below yield are small, and elastic distortion in axial loading is usually not visible to the naked eye. For a typical hot-rolled, mediumcarbon steel, 55 to 60 ksi (379 to 414 MPa) is a reasonable value for the yield strength and the modulus of elasticity is about 30 106 psi. Using Equation 8, the elastic strain at the onset of yielding is about 0.002 in./in. Permanent plastic strain at fracture is much larger in a ductile material and

26 Types of Failure and Stress

is usually visible without magnification. It can vary from very little (less than 5%) for an age-hardened alloy or lightly tempered steel to as much as 40 to 50% elongation for a soft annealed alloy. Now assume that the load is increased to P2, and again the load removed. When the specimen is unloaded, a portion of the length is recovered, but the unloaded final length is greater than the original length. The recoverable portion of the length is elastic deformation, and we describe the unrecovered length change as being plastic deformation. The stress at which permanent deformation begins is the yield strength of the material (usually measured when the plastic strain is 0.2% as indicated by the line in Figure 2). A second potential failure mechanism has now been identified permanent deformation. If failure is due to plastic yielding, the solution is to either increase the cross section to decrease the stress or select a material with a higher yield strength. Yield strength is one mechanical property strongly affected by the microstructure so that two material solutions are possible: select a different alloy of higher yield strength or mechanically/thermally process the same alloy to obtain a higher yield strength. This might be done by selecting a cold-worked grade of the material if the failed component was in the annealed or hot-rolled condition. Alternatively, specifications for the part may include a maximum permissible grain size because the yield strength of the material varies inversely with the square root of the grain diameter (d) and constant (k) according to:
Sy k d (Equation 9)

As will be seen in following lessons, grain size is an important microstructural variable that affects many properties in addition to yield strength. If the load is still further increased beyond the maximum (Pu), the specimen shows a continued increase in length, but the cross-sectional area of the specimen now varies with position along the length, and we say that a neck has developed (Figure 14). Necking is evidence of gross overload, and fracture occurs within the necked region.

Figure 14. Necking and the cup-and-cone fracture that occurs in tensile loading of a ductile material (a) and flat fracture associated with brittle behavior (b).

Failure in Tensile Loading

27

The maximum load obtained in the test divided by the original cross section is the tensile strength, ultimate strength, or ultimate tensile strength (commonly abbreviated as UTS) of the material. Additional causes for failure are associated with this deformation: The total strain to fracture may be too small, and the material would then be described as (macroscale) brittle. The relative ductility of the material is, like the yield strength, strongly affected by the microstructure; so an improper heat treatment might be the cause for failure. The difference between the yield strength and the tensile strength is a margin of safety against fracture. If a load incursion above the design stress occurs when the yield-totensile ratio is low, the result is simply increased plastic deformation. An improperly designed spring would fail to operate properly if permanently distorted, but no fracture has occurred, and the load can be removed from the component without fracture. At higher loads, deformation becomes localized by neck formation and a cup-and-cone fracture occurs in axial loading (Figure 14). If a load incursion occurs in a structural member when the yield-totensile ratio is high, the load incursion may not only cause the yield strength to be exceeded, but also the tensile strength. In such an instance, ductile fracture may occur at the microscale but not be visible at the macroscale. The difference between the yield and tensile strength is due to the strain-hardening capability of the material, and the strain at the onset of necking correlates with the amount of strain hardening. For low strainhardening materials, the YS/UTS ratio is high, and, for high strain-hardening materials, the ratio is low. There is a general trend for strain hardening to decrease as the strength level increases. Steels quenched and tempered to high hardness, cold-worked materials, and age-hardened alloys have low strain-hardening capacity. The YS/UTS ratio may also be high because the material has inherently low ductility. In this case, a load incursion above the yield strength may also result in load incursion above the ultimate strength, but there is no ability to deform plastically. The elastic strain energy stored in the specimen under load cannot be dissipated by plastic deformation, and it is then available to drive a brittle crack through the material. This is a dangerous situation because there is no prior warning via plastic deformation that fracture is imminent, and the crack propagates faster than the load can be removed to stop crack propagation. The previous discussion would indicate that safety in design against plastic deformation is achieved by increasing the load-bearing cross section so as to decrease the stress. This may not be the case. The presence of an imperfection in a thin component may result in unexpected fracture, but the fracture is by a ductile mechanism if the material is inherently ductile (see Lesson 3). This in turn implies that the crack will stop propagating unless the load continues to increase. The presence of an imperfection in a heavy section may cause an inherently ductile material to fail in a brittle way and, therefore, may change the fracture mechanism to one which occurs with no warning. This is easily observed by taking the same material as in Figure 2 and introducing a stress concentrator such as a Vgroove into the specimen. The load-elongation curve now changes as

28 Types of Failure and Stress

shown in Figure 15, which indicates that there is no plastic deformation prior to fracture and, therefore, no warning that fracture will occur if the yield strength is exceeded. Because there is no strain hardening to absorb the stored strain energy and because the crack propagates very fast, there is no way to decrease the load to prevent crack propagation. If failure occurs due either to distortion or fracture, it is critical to remember that fracture initiates in a region where the internal stresses created by the external loads and the component geometry exceed the strength of the material. In the case of fracture, the load-carrying capability may be described by a fracture stress or by a critical value of the stressintensity factor, K (Lesson 3). In the case of plastic distortion by extension or compression, the governing material property is the yield stress. Variations in local stress in the component may be due to the way in which the load is applied, for example in bending and torsion, and/or the presence of microscale stress raisers. Alternatively, the stress may vary due to the presence of deliberate changes in geometry such as the presence of threads, keyways, oil grooves, and so forth or inadvertently due to imperfections in the material. Examples of material imperfections include surface pitting due to corrosion, a partial lack of fusion in a weldment, quench cracks produced by heat treatment, and primary manufacturing imperfections such as shrinkage porosity during solidification and laps and seams created during forming operations. The strength of the material is in turn controlled by the microstructure, which may also vary with location in the component. Again, the variation may be deliberate or inadvertent. Deliberate variations in strength occur, for example, in surface-treated components where the material is heat treated to produce a high-strength and wear-resistant case supported by a lowerstrength and tougher core. Typical processes include induction hardening,

Figure 15. Loss in ductility in a smooth section tensile specimen as a notch is introduced. Stress at the onset of yield is increased, fracture stress increases, ductility decreases, and work to cause fracture decreases (area under curve).

Failure Induced by Compressive Loading 29

carburizing, and nitriding. Splines, gears, and shafting are typical examples of surface-treated components. Inadvertent variations in material strength may occur for example by decarburization of steels during heat treatment. Whether decarburization degrades the load-carrying capability of the material in turn depends upon the type of loading and the internal stresses that are created by the loads. Axial loading of a slightly decarburized material (for example, 10 mils, or 250 m, on a 0.5 in., or 12.7 mm, diam shaft) does not cause significant degradation in load-carrying capability in axial loading relative to that caused by bending or torsional loads. At first impression, failure would seem to be less likely in compression. Because the load-bearing cross section increases during deformation, necking cannot occur in ductile materials. Also, it is difficult to imagine crack propagation on a plane subjected to a compressive stress because the stress will close the crack. Compressive loading can cause secondary tensile stresses in the material, and it is these stresses that may cause crack propagation and fracture. Possible failure modes then include distortion that may result in buckling as well as fracture. Excessive deformation may occur due to the same factors cited in the case of tensile loadingtoo low a yield strength, too low a modulus, too small a section size (stable deformation). However, a new failure mode identified as buckling, which is an unstable deformation process, can also occur. Ductile fracture may occur in three different ways: (1) by nonuniform deformation through the height of the component and crack initiation at the surface on a hoop plane, (2) line loading crack growth and/or initiation and growth at the centerline, and (3) by shear band formation. Brittle fracture can occur on a plane of maximum shear stress and on a plane of maximum tensile stress. Each of these failure conditions is discussed in the following section. Small Height-to-Diameter Ratio. An important variable in controlling which of the failure modes described previously operates is the height-todiameter ratio of the component. Assume for the moment that the load is distributed uniformly across the cross section, say by the platen shown in Figure 16, and that the height-to-diameter ratio is small (say less than 2). Also assume that the material is inherently ductile. Because the material wants to expand perpendicular to the load, there is friction between the platen and component. This friction can restrict material adjacent to the platen from moving laterally, and it often results in bulging of the specimen at half height. If the material bulges at half height, the increased circumferential strain in this region creates an internal normal stress on circumferential planes that is largest at the outside diameter. Horizontal sectioning of the specimen reveals cracking on the hoop planes, and these cracks may be visible on the external surface depending upon their size and the roughness of the surface (Figure 16). The cracks start at the outside surface and grow inward as the load is increased. Increased surface roughness and/or draw marks from previous working increase the likelihood of this cracking mechanism. Cracking may also initiate near the centerline of the component in two cases (Figure 17). (1) Some fabrication processes create centerline

Failure Induced by Compressive Loading


Stable Deformation and Unstable Deformation (Buckling)

Height-to-Diameter Effects

30 Types of Failure and Stress

Figure 16. Compression of a plate. Friction forces (F), which increase as the compressive force P increases, at the platen/specimen interface cause restricted lateral flow near the platen, but no restriction away from the platen. Bulging at midheight causes circumferential strain in the bulge. Cracking can occur in the bulged area. The cracks start at the surface on the circumferential planes and grow inward as height reduction increases. imperfections, for example swaging (Figure 18) and forging. As the material is compressed during forging, it tries to expand laterally, and crack growth may initiate at the centerline defect, growing vertically in the figure. (2) Even in the absence of a pre-existing centerline defect, crack initiation can occur along the centerline and spread vertically for line loading. This becomes more likely as the ductility of the material is decreased. Brittle fracture in compression can occur in two ways: (1) vertical splitting along the centerline as indicated previously and (2) on a plane of maximum shear stress that is at 45 to the applied load. The latter is a common fracture mode for cast irons. If a small height-to-diameter ductile specimen is loaded in compression to some fairly large permanent reduction and then sectioned and macroetched, there is evidence that strain has concentrated along (shear) bands in the material (Figure 19). Sectioning after further compression shows that crack initiation can occur at two locations: at the componentplaten interface or at the intersection of two shear bands. Large Height-to-Diameter Ratio. When components are tall with respect to their diameter, friction forces are minimized. Additionally, it is

Figure 17. Crack locations in compression loading. Source: Ref 6.

Failure Induced by Compressive Loading 31

Figure 18. Centerline cracking in a swaged rod. difficult to apply the load totally uniformly along the axis, which creates small bending moments in the material. This loading condition causes failure to occur by buckling. Buckling of a column loaded in compression is shown schematically in Figure 20. Buckling is unstable deformation in that once buckling initiates the load required to increase the buckling deformation decreases. Buckling may

Column Buckling

Figure 19. Shear bands and crack initiation within a shear band for material loaded in compression. Source: Ref 7.

32 Types of Failure and Stress

Figure 20. Column bucking for different connections. See text for discussion. Source: Ref 8. be initiated at stresses below the yield strength (elastic buckling) or above the yield strength (plastic buckling). Whenever compressive loads are present, buckling competes with excessive uniform deformation or brittle fracture as a failure mode. Buckling may also occur due to twisting loads (torsion). Buckling may be initiated by loads creating a stress less than that corresponding to net-section plastic deformation and, therefore, at a stress less than the net-section yield strength (elastic buckling). In some cases, structural members are designed so as to prevent elastic buckling, but buckling may still occur at a stress greater than the yield strength (plastic buckling, Figure 21). A structure in which a member buckles elastically in service may be restored to good operating condition by releasing the load on the member. Elastic buckling may be avoided by a redesign in which the load distribution is changed. However, a member that buckles due to plastic loading has failed and must be replaced. In an inherently ductile material, the load required to increase the amount of bending decreases as the bend-

Figure 21. Compression plastic buckling in compression loading of a tube. Source: Ref 3, p 374.

Failure by Torsional Loading

33

ing increases, so that, if the load is increased, elastic buckling progresses to plastic buckling followed by collapse and ultimately fracture. The critical load for column collapse depends on the slenderness ratio (length divided by cross section) of the column as well as the modulus of elasticity of the material and to some extent on how the column is affixed to other members. The latter is taken into account with the K factor, shown in Figure 20. K factors range from 2.0 for a column free at one end and fixed at the other end to 0.5 for a column rigidly fixed at both ends. Columns that are long and slender and that have a small modulus of elasticity have the smallest critical loads. Columns that are rigidly fixed on both ends (K 0.5) have higher collapse loads than do columns fixed only at one end (K 2 if the other end is free, K 1 if other end is pinned). Because the modulus of elasticity does not vary with microstructure (and therefore not with heat treatment) and only slightly with composition, buckling resistance is essentially determined when the component is sized if the material has been selected. Buckling may occur in beams subjected to transverse loads. Loading a beam in bending creates a compressive stress on one side of the neutral axis, parallel to the length of the beam, and it is this stress that can cause buckling. Again, as for axial compressive loading, buckling depends on the modulus of elasticity of the material and the slenderness (in this case, beam web height compared to the web width and flange height to flange width). Deep and thin web sections and wide and thin flanges as in H- and I-section beams are more susceptible to buckling. The onset of brittle fracture comes without any warning because there is no indication of prior plastic deformation to suggest the onset of impending failure. In the same way, buckling failures sometimes provide no or little visual warning. An example is snow loading on a roof. Elastic deformation (sagging) of the roof increases as snow loading increases. It is not customary to measure the roof sagging, so the roof may be heavily loaded, and, if more snow accumulates, the buckling load is exceeded and failure is catastrophic because collapse can continue with continued decreasing load. It would have been possible conceptually to determine that failure was imminent, but, from a practical standpoint, this is not done, and roofs are supposed to be designed to resist buckling failure. For any loading condition in which compressive or torsional stresses are present, buckling must be considered as an alternative failure mode. Just as for tensile and compressive loading, deformation prior to fracture may be uniform or nonuniform. Excessive distortion failures may occur for the same reasons as for axial loading: too small a shear modulus, too low a yield strength, or too small a cross section. Additionally, buckling can occur in torsion of hollow tubes. Again, the critical buckling load depends on the geometry of the tube and the elastic (shear) modulus. The collapse load increases as the modulus increases and decreases with the square of the tube length. Thin-wall tubing is especially prone to buckling failure as opposed to uniform deformation.

Failure by Torsional Loading


Excessive Deformation

34 Types of Failure and Stress

Ductile and Brittle Fracture

Torsional loading differs from axial loading and bending loading in that the ratio of maximum shear to normal stress is 1 to 1, whereas in axial and bending loading it is 1 to 2. Cracking is now equally likely on shear planes and planes of maximum normal stress. Any plane along the length of the member perpendicular to the axis or any longitudinal plane containing the axis is a plane of maximum shear stress. Consequently, cracking could initiate at any location along the length on either of these planes. Any plane at 45 to the axis as in Figure 7 is a plane of maximum normal stress, and again there are an infinite number of these planes intersecting the surface along the length. A macroscale brittle fracture results in a helical fracture as shown in Figure 22. A ductile fracture surface is shown in Figure 23. Close examination of this fracture surface often reveals secondary cracks on longitudinal shear planes on which the stress is the same as on the transverse planes. Because the ratio of the stresses is the same on shear and normal planes, it is not uncommon to find multiple crack initiation sites that start on both planes of maximum normal stress and maximum shear stress, especially under cyclic loading. As these cracks grow, coalescence occurs and the fracture surface can take on a star pattern. Crack initiation induced by a machining mark in monotonic loading can be on a shear plane and then turn and run on a normal stress plane. Alternatively, draw marks on a forged or drawn shaft can cause stress concentration and crack initiation on a plane of maximum shear stress, but the crack then turns and runs on a helical plane of maximum normal stress. The ratio of maximum shear stress to maximum normal stress in bending is 1 to 2. The maximum nominal tensile stress occurs where the maximum moment occurs (assuming constant section modulus) and at

Failure in Bending Loading

Figure 22. A truck rear-axle side shaft that broke during service. The fracture is a torsional fatigue fracture originating at a grinding mark (shown at arrows). The fracture surface is helical and inclinced at an angle of 45 to the axis of the shaft. Left view shows the fracture with fatigue cracks visible. The right view shows the grinding mark on the surface at fracture origin.

Failure in Bending Loading

35

Figure 23. Macroscale ductile torsion fracture. Fracture is on a plane of maximum shear stress (perpendicular to the longitudinal axis). Examination at low magnification may also reveal so-called secondary cracking on other planes of maximum shear stress parallel to the longitudinal axis. the maximum distance from the neutral axis. It may be necessary to perform a stress analysis to determine this location for complex geometries. A few simple cases are shown here. In a three-point, symmetrically loaded beam or in a uniformly loaded beam, the maximum moment is at half the span length (Figure 24). In a four-point, symmetrically loaded beam, it can be at any place between the two loads because the moment is constant between these two locations(Equation 1). The bending moment varies more slowly along the length for a distributed load so that the expected site of crack initiation is wider. A brittle crack is expected to start on a plane perpendicular to the axis of the beam, at the location of maximum tensile stress. A ductile crack in a homogenous material would start on a plane at 45 to the plane of maximum tensile stress. This rotation of planes is around a direction parallel to the width of the beam (Figure 25). A ductile crack could theoretically initiate due to the shear stress at the neutral axis on a plane parallel to the axis of the beam. However, the ratio of the maximum value of this shear stress to the maximum normal

Figure 24. Bending moment distribution along a beam. (a) Three-point loaded beam (concentrated load at L/2). Mmax PL/4. (b) Uniformly distributed load of (w) pounds/length. Mmax wL2/8. If the total force in (a) is distributed, the maximum moment is half that in (a). (c) Four-point loaded beam. Mmax Q(x/L).

36 Types of Failure and Stress

Figure 25. Planes of macroscale ductile and brittle crack initiation for a three-point loaded beam. See text for discussion. stress is small. For a three-point loaded beam it is 4/9 (h/L; h is height, L is length of beam). Similarly, brittle fracture could be initiated at the neutral axis on a plane rotated 45 from the plane of maximum shear stress. This is unlikely for the same reason. However, if the material is not homogeneous at the neutral axissay a soldered or welded joint, failure could initiate at the neutral axis, especially if the area of the solder or weld is too small. Macroscale brittle crack initiation on a transverse plane at the location of maximum bending stress is common. However, as the crack propagates, it often curves toward the longitudinal axis of the component (Figure 26), and microscale examination shows ductile fracture in the curved portion of the fracture surface. Buckling failures may also occur in beams. Consider a symmetrical Hbeam. The compressive longitudinal stress acting in the beam can cause buckling of either the flange or the web due to insufficient thickness.

Summary of Excessive Deformation Failure Modes

This section of the lesson summarizes the potential failure modes that result in excessive deformation. Distortion failures may be quasi-stable in the sense that macroscale distortion is present in a component, but collapse has not occurred. In some cases, distortion may not be macroscale visible. If it is due to deadweight loading, an increase in load may cause sudden collapse. In that sense, failure is just as unanticipated as it is in the case of brittle fracture. The macroscale appearance of a buckling failure is obvious.

Summary of Excessive Deformation Failure Modes 37

Figure 26. Fracture due to bending in a cylindrical section (SAE 1050 axle shaft) at location 2. Crack initiation occurred on a plane of maximum normal stress at the surface of the section. However, as the crack propagated, the fracture plane curved toward the longitudinal axis. (Note also that the component has been surface treated as evidenced by the smooth outside circumferential fracture zone and the rough interior zone.) Four types of failure by distortion are recognized:

Yielding. The stress created by the applied loads exceeds the yield strength of the material in tension, compression, torsion, bending, and combinations of these loading conditions. The solution is to choose a material of higher yield strength or to increase the section thickness. An example is the loss of operation of a spring due to excessive stretching, bending, or twisting if the spring has not been preset. Inadequate Stiffness. The elastic moduli (tensile modulus and shear modulus) are too small. Unfortunately, the elastic moduli of metallic materials cannot be changed by heat treatment, and, furthermore, there is little variation of the elastic modulus among the common metallic alloys (a factor of about three). Consequently, distortion can be decreased only by increasing the section thickness and/or by decreasing the load. Distortion of an axially loaded member is decreased by reducing the stress, which in turn is accomplished by increasing the cross section. The stiffness of a beam loaded in bending or of a section loaded in torsion can be increased if the section modulus is increased. (Other types of load instability can occur in compressive and torsional loading besides buckling and necking. They are discussed in Lesson 3.) The modulus of materials decreases as the temperature is increased. For elevated-temperature service, it may be possible to identify a substitute material having a higher melting point and, therefore, a higher modulus at the operating temperature. Failure due to Buckling. Buckling is instability that occurs due to compressive or torsional stresses, usually of a long slender part or

38 Types of Failure and Stress

region of a component. The stability of the component depends on the slenderness ratio and the elastic modulus. There is a factor of about 3 between the stiffest and least-stiff metallic materials. In a practical situation, the stiffness usually cannot be changed by changing materials because of other design considerations (corrosion, wear, etc.), but it may be increased by appropriate alteration of the geometry of the cross section.

Creep/Stress Relaxation. Time-dependent strain at a homologous temperature greater than about 0.4. The metal gradually changes its shape and dimensions over a relatively long period of time. The environmental service conditions fix the temperature, and, therefore, little can be done to control creep deformation. Selecting a material of higher melting point can improve the loss in modulus at elevated temperature. Creep rates can be controlled to some extent by alloy and microstructure selection. Scaling resistance can be controlled by alloy selection. When a component is clamped under a tensile load at elevated temperature, it is observed that the clamping load is lost after some time due to stress relaxation; that is, the load required to maintain a given deformation decreases with time at elevated temperature. This is an important failure mechanism in bolted assemblies subjected to elevated temperature.

States of Stress in Common Geometries

Macroscopically, there are two types of fracture, ductile and brittle. It is assumed that if fracture occurs on a plane of maximum shear stress it is ductile, and, if it occurs on a plane of maximum normal stress it is brittle. Therefore, it is essential to know the locations of the maximum normal and shear stresses in a manufactured component. The lesson to this point has discussed the state of stress in axial loading (tension and compression), bending, and torsional loading of cylindrical sections. There are other common geometries of manufactured components including curved beams, eye bars, hooks, and so forth. The state of stress in these components is not discussed, but note that there are standard reference texts such as most mechanical engineering machine design texts that discuss these geometries. Another good source of this kind of information is Blake (Ref 9). Two additional geometries discussed in this section are thin-walled pressure vessels and manufactured components subjected to direct shear loading. Consider the internally pressurized cylindrical section with end caps shown in Figure 27, having a radius r and a wall thickness t. Assume that r/t is large, for example, 20. Now section the cylindrical portion as shown. It is clear that a normal stress must exist in the circumferential direction across the thickness of the wall. Similarly, if the cylindrical section is cut perpendicular to the axis of the cylinder, there must be an axial stress in the section. Writing a force balance in each of these cases, we determine that the hoop stress is approximately twice the axial stress, that is,
Shoop Saxial pr t pr 2t (Equation 10)

Thin-Walled Pressure Vessels and Piping

(Equation 11)

States of Stress in Common Geometries 39

Figure 27. Stresses in an internally pressurized thin-wall section. where p is internal pressure, r is vessel radius, t is wall thickness, and so long as the wall is thin, that is, r/t greater than 10, any shear stress in the wall and the presence of a radial stress may be neglected. Based on the magnitude of the above stresses, we would expect brittle fracture to occur on a plane perpendicular to the hoop stress. If ductile fracture occurs, it results in fracture on a shear plane running the length of the vessel and rotated 45 around the length direction from the hoop plane. This leads to a characteristic appearance described as a fishmouth fracture (Figure 28). For piping, the allowable pressure may be based on the magnitude of the axial stress. Specifications sometimes use a more accurate expression than the approximate equation (Eq 11)(Ref 4), e.g., ASTM B 88 for copper tubing requires the axial stress to be calculated according to:
Saxial p(2r 0.8t) 2t (Equation 12)

The vessel must be supported in some way. Therefore, we should also consider the effect of any bending stresses that might be present. For thinwalled vessels, the weight of the tank or pipe does not contribute a very large bending stress to the loading condition and can therefore often be neglected. The above analysis omitted the weight of the material inside the vessel and is therefore representative of tanks containing pressurized gases. Consider the situation where the weight of the tank can still be neglected, but the weight of the material in the tank (the volume of the material times its density, say water) cannot be neglected. Any bending stresses created by the weight of the tank plus its contents algebraically add to the axial stress created by the pressure. Therefore, the fracture location is predicted to change from any hoop plane around the circumference to a location on the bottom of the tank at maximum bending moment and on an axial plane. Analysis shows that this will be the fracture location if the bending stress in the vessel exceeds the axial stress created by the pressure.

40 Types of Failure and Stress

Figure 28. Fishmouth fracture in a thin wall internally pressurized section. (a) Overall view. 0.5. (b) Unetched section from location between arrows in (a) showing extensive transverse cracking adjacent to the main fracture (at right). Approximately 4.5. (c) Specimen etched electrolytically in 60% HNO3 (nitric acid) showing intergranular nature of cracking. 100. Internally pressurized piping contains the same axial and hoop stresses created by the internal pressure and the bending stresses created by the distance between supports. However, no axial stress is created when the pipe ends are open (not capped or plugged).

Direct Shear Loading

Many coupling and locking devices are subjected to a state of stress described as single or double direct shear loading. Two examples are shown in Figure 29. The top figure is described as single shear, and the bottom figure as double shear loading. Single shear loading is analogous to cutting a piece of paper with a pair of scissors or the action of a shear blade on a rolling mill line to cut off rolled lengths of material. Single shear loading can also occur on locking pins and keys in a shaft. The average shear stress in the pin in this condition is the shear force divided by the cross section:
avg P A (Equation 13)

Now consider a clevis containing a pin. Loading on the pin is double shear. Failure could occur in the pin by crushing, or by shear fracture.

States of Stress in Common Geometries 41

Figure 29. Single shear loading on a pinned connection (top). Double shear loading in a clevis connection (bottom). Shear fracture in the pin occurs on two areas (assuming that the load is uniformly distributed across the pin). Thus, the average stress is now given by:
avg P 2A (Equation 14)

With reference to Figure 30, other failure modes are possible. (1) The force P may cause crushing in the plate on the compression side of the hole. (2) Alternatively, if there is an inadequate amount of material above the hole in the plate, tear-out of this plug of material can occur by shear. A common rule of thumb to prevent this failure mode is that the distance y in Figure 30 should be in the range of 1.25 to 2 times the pin diameter. (3) There is stress concentration in the hole at the 3:00 and 9:00 positions. Fracture can initiate at this location and propagate transversely to the edge of the plate. A careful reading of this section shows that the fracture surface orientation can be correlated with the type of loading. This correlation is an integral part of the failure analysis. In the same way, we have introduced the concept that both stress and strength can vary across the cross section of a

Conclusions

42 Types of Failure and Stress

Figure 30. Shear loading on a pin or rivet connection. In order to prevent shear failure in the plate, the hole is usually located so that the distance y is between 1.25 and 2 times the hole diameter. machine component. Fracture starts where the stress created by the external load and by the part geometry (local stress) exceeds the local strength of the material. Therefore, it is essential that we know if strength varies across the cross section (identified by microstructural analysis and/or microscale hardness testing) and the location of maximum local stress. The latter requires that either the analyst has the background to do nominal stress analysis and recognizes geometric material imperfections or that the analysis is done by someone else and incorporated into the study of the failure. Standard reference books, sometimes devoted to specific geometries (e.g., pressure vessels), are available. See the References and Selected References at the end of this lesson.

ElevatedTemperature Environments

There are a large number of environmental design considerations associated with both manufacturing operations and service environment, many of which are alloy specific. The following list is not complete, but it does indicate the scope of consideration. Pickup of gaseous hydrogen may result in premature failure due to embrittlement, as can radiation damage and liquid metal embrittlement. Mercury cracking of brass is one example of liquid metal embrittlement. Carbon pickup can result in the presence of deleterious carbides in some materials and elemental flake graphite in other instances. In either situation, local ductility is lost. The stability of the microstructure must be considered if the service temperature becomes high. An example is in situ graphitization resulting from the decomposition of carbides in steels of low alloy content (and either high molybdenum or aluminum content). Long-term, elevated service temperatures may also result in the formation of new carbide phases that have a different morphology from the parent carbide and may also have a different location in the microstructure. The formation of sigma

Residual Stresses 43

phase resulting in embrittlement in high-chromium steels is a specific example. A second example is the sensitization of austenitic stainless steels caused by the precipitation of chromium-rich carbides in the grain boundaries and concurrent depletion of chromium in the adjacent matrix. Overaging of precipitation-hardening alloys, often accompanied by the presence of coarse grain boundary phases, also results in a loss of strength and toughness. These environmental considerations are examined more completely and in more detail in subsequent lessons. However, it should be clear from this abbreviated listing that, in many instances, degradation occurs on the microscopic scale. This is turn emphasizes the importance of macrostructural and microstructural examination of appropriate specimens removed from the component as part of the failure analysis. Residual stresses are locked-in elastic stresses that exist in a component due to mechanical and thermal processing during manufacture. Once created, they remain in the component unless a subsequent manufacturing operation causes their relaxation or the part distorts or cracks. They can therefore be present in a component when it is placed in service. Essentially all manufacturing operations that involve plastic deformation and/or thermal treating potentially create residual stresses; that is, they are a natural consequence of mechanical and/or thermal processing. Residual stresses can be either degrading or beneficial to subsequent service performance. Some manufacturing operations deliberately introduce residual stresses in the component to improve service performance. In other situations, a manufacturing operation creates harmful residual stresses, and, in critical applications, additional processing is required to mitigate the harmful effects of these stresses. Harmful effects may occur during any manufacturing step following the one in which the stresses are created, or their effect may not be seen until the finished component is placed in service. Residual stresses can cause distortion of a part, reduced load-carrying capability, and, in extreme cases, can cause cracking without the presence of an additional external load (quench cracking, hydrogen embrittlement). Sections that are thin in one dimension can warp excessively due to the presence of residual stresses. Residual stresses contained in a component after completion of the manufacturing operations algebraically add to the stresses created by the service loading conditions and can either increase or decrease net stresses in the component. If surface residual compressive stresses can be created in a part subjected to a bending load in service, the net stress on the tensile side of the bend created by the applied load is reduced by the amount of the longitudinal residual compressive stress. Unfortunately, some manufacturing operations introduce tensile residual stresses at the surface of the component and, therefore, degrade service performance. Machining operations introduce tensile residual stresses, and grinding can introduce either compressive or tensile residual stresses. Abusive grinding is assumed to create tensile surface stresses. Quenching operations often produce beneficial surface compressive residual stresses. However, equilibrium requires that compressive stresses in one region be balanced by tensile

Residual Stresses

44 Types of Failure and Stress

stresses in another region. Then, tensile stresses created in bending may add to subsurface residual tensile stresses, causing crack initiation below the surface of the material. Processes designed to impart beneficial residual stresses are especially important to the improvement of service performance of components loaded in long-life fatigue, as well as for components sensitive to stresscorrosion cracking and corrosion fatigue. These processes include thermal surface treating, shot peening, mechanical prestressing, surface rolling, and burnishing. The consequences of harmful residual stresses include spontaneous fracture or, in other instances, delayed cracking. Some parts such as structural beams and large rollers (rolls in a rolling mill) have fractured spontaneously without application of any external load. Some common types of cracking caused by residual stresses include grinding cracks, heat checking, quench cracking, weld cracking in either the heat-affected zone or in the hot weld metal, and thermal fatigue. In addition, distortion associated with grinding, welding, and heat treating is caused by residual stresses. There are standard procedures to determine experimentally the magnitude and distribution of residual stresses in components. They include nondestructive x-ray techniques and destructive, material-removal procedures. However, these procedures are time consuming and expensive, so, in some industries, they are not carried out unless a special and expensive problem develops, for example a change in a manufacturing operation that has resulted in a large number of field failures. However, the distribution of residual stresses created by manufacturing operations must at least be understood qualitatively so that the possibility of introducing harmful residual stresses can be included in the design of a manufacturing process and in failure analysis. Residual stresses are locked-in elastic stresses that exist inside a component. They are created when the manufacturing operation creates a strain gradient inside the component that is not subsequently relieved. Strain gradients can be created in four ways:

Nonuniform temperature in a part during either heating or cooling Nonuniform plastic deformation during cold working When composite materials having different elastic moduli and coefficients of thermal expansion are mechanically or thermally loaded Chemical deposition processes (including electroplating)

When a machine component is placed in a furnace at elevated temperature, it does not reach the furnace temperature instantaneously. The outside of the component reaches the furnace temperature before the interior does. Therefore, a thermal gradient is created in the component during heating. A similar nonuniform temperature gradient is created when the component is removed from the furnace. Because almost all materials

Residual Stresses 45

expand when heated and contract when cooled (in the absence of a phase transformation), differences in temperature imply that the hot part of the material is larger than the cooler part. Unless the component breaks due to this difference in volume, the material must accommodate the difference by straining the larger volume region in compression and the smaller volume region in tension. These strains then create elastic stresses (stress strain times modulus). A helpful analogy to understand how residual stresses are created is by the use of a spring analogy. Assemblies such as those shown in Figures 31 and 32 created with tension springs (or rubber bands) and compression springs (or rubber pads) help to explain the concept. Springs in tension in the assembly represent the presence of tensile residual stresses (the springs are longer than their no-load length and therefore contain a restoring force) and springs in compression represent the presence of compressive residual stresses (the springs are shorter than their no-load length). The spring-and-plate assembly in Figure 31 illustrates the case of two plates connected with three springs, one of which is adjustable. There is no stress in the system if all springs initially have the same length. However, if the springs are initially of different lengths when the component is assembled, stresses are created in the springs because all springs are forced to have the same length. If the adjustable spring is initially shorter than the two fixed springs, putting the system together requires that the center spring be lengthened and the two outside springs shortened. Therefore, there are restoring forces or locked-in stresses in the springs, producing the state of stress shown to the right (surface compression, centerline tension). The same experiment can be performed in which the adjustable spring in the center is longer than the fixed-length springs (Figure 32). Now, the residual stress pattern is reversed from that in Figure 31 with residual tension at the surface and compression in the interior. Several things can be learned from the two models in Figures 31 and 32:

Visualizing Residual Stresses

Each system is in static equilibrium. If any spring is removed, the system will collapse; that is, the distance between the plates will change.

Figure 31. Three-spring assembly resulting in residual stresses. The initially short center spring is stretched with an adjustable screw. The state of stress is then tension in the center spring and compression in the two outside springs.

46 Types of Failure and Stress

Figure 32. Three-spring assembly resulting in residual stresses. This case is the opposite of that shown in Figure 31: tension in the two outside springs and compression in the center spring.

The vertical force in a set of springs is balanced by an equal and opposite force in another spring. That is, it is the force that is balanced, not the stress. The difference between the force in each spring and the stress in each spring is fixed by the spring constant of each spring, that is, the stiffness of the spring. Each spring is stressed so that the tension springs are longer and the compression springs shorter than their load-free lengths. Each spring loaded in tension would like to be shorter, and each spring loaded in compression would like to be longer to relieve the internal stress. If the length of any spring is changed, the system will change shape (distort) to reach a new static equilibrium position.

Equilibrium of the system requires that the positive spring forces be equal to the negative spring forces. Figure 33 shows two residual stress patterns created by shot peening and a superimposed bending stress in both cases. Considering only the residual stresses, the vertical distance represents the magnitude of the stress and horizontal distances represent the area in which the stress exists. Therefore, because force stress area, the total tensile area must equal the total compressive area. As is discussed later in this section, it is desirable to have surface compressive residual stresses in a component subjected to bending. Therefore, if the surface compressive stresses are of large magnitude and too deep, the subsurface tensile stress increases in magnitude. If the tensile stress becomes too large, say in bending, crack initiation can occur below the surface because the tensile residual stress adds to the stress created by the applied load. With the spring analogy in mind, the relation to actual assembled components can be visualized. For example, consider a bar containing the stress distribution shown in Figure 30(left): surface compressive residual stress and interior tensile stress. If a hole is drilled along the centerline of the bar, the tensile residual stress disappears, and the compressive stresses on the surface now relax, causing the bar to increase in length. The residual stresses can be relaxed in a different way. Assume that the exterior material is removed by machining. Now the material along the centerline containing tensile residual stress can relax because it is no longer constrained by the outer compressed material. The result is that the now

Residual Stresses 47

Tension

Compression

Bending stress

Net stress

Figure 33. Residual stresses from shot peening and a superimposed bending stress. The magnitude of the surface residual stress and the area over which it exists must be balanced by the magnitude of the internal stress. That is, the total tensile stress area shown in the figure (stress times depth in the figure) must be balanced by the total compressive stress area. Large deep compressive stresses on the surface increase the magnitude of the subsurface tensile stress. When this assembly is loaded in bending (or torsion), the net stress is the sum of the bending stress plus the residual stress. This can result in subsurface fracture in the region of residual tensile stress. smaller diameter bar decreases in length. The general observation about changes in length and resultant residual stresses can be obtained from this discussion. Material that is shorter than it would like to be ends up in residual tension when the residual stresses are created. Material that is longer than it would like to be ends up in residual compression. Alternatively, material containing residual compressive stresses increases in length when the load is removed and vice versa. In this example, the residual stress pattern was symmetric about the axis of the bar. Other examples are examined for which symmetry does not exist, causing members to warp or bend. The actual state of residual stress in manufactured components of complex geometry is considerably more complicated than that in the example above. The stress state is three dimensional, rather than one dimensional. For a cylinder, we consider residual stresses in the axial, circumferential,

Components of Residual Stress

48 Types of Failure and Stress

and radial directions. For a flat plate, we should consider residual stresses in the length, width, and thickness directions. In many instances of plate and cylinder geometries, it is usually possible to eliminate consideration of the stress in one direction. For cylindrical geometries, the radial direction can often be neglected, and, in plate and thin-disk geometries, the thickness direction can often be neglected. It is not always possible to neglect the stress in some direction when the component is a complicated cast shape.

The Bicycle Wheel and Wagon Wheel

Consider a bicycle wheel (Figure 34a) and a wagon wheel (Figure 34b). The spokes of the bicycle wheel are tightened, producing tension. This in turn draws the wheel rim inward, creating circumferential compression, and the wheel hub outward (tension). The axial stress along the centerline of the hub is negligible. A wagon wheel is fabricated of wood components held together by a steel rim. The wooden part of the rim normally is made of multiple pieces, and the spokes and hub are also fabricated of wood. The steel tire is carefully sized to be smaller in diameter (and therefore circumference), than the wooden part of the assembly. The steel rim is heated so that it expands (increases in circumference), and it is then placed around the wooden assembly and allowed to cool. As it cools, it shrinks, decreasing in circumference, which loads the wooden part of the rim in compression as well as the radial wooden spokes that in turn squeeze the hub in compression. Therefore, all wooden parts are loaded in compression. The balancing circumferential tensile forces are in the steel rim.

Measurement of Residual Stresses

Standard procedures for the measurement of residual stresses fall into two groupsdestructive (or dissection techniques) and nondestructive. The former procedure requires removal of metal, either by drilling out material or by machining material from the surface (sectioning techniques). Strain gages are usually applied to measure the changes in shape, and the stress state is determined from the strain measurements. Difficulties can

Figure 34. (a) Bicycle wheel. Tensile stresses in the spokes create circumferential compressive stresses in the rim. See text for discussion. (b) Wooden wagon wheel. The steel rim is heated and placed around the wooden rim sections. When the heated steel cools and shrinks, it creates compressive hoop stresses in the wooden rim.

Residual Stresses 49

arise in these metal-removal procedures when the stress distribution varies through the thickness; that is, the stresses vary in three dimensions, not simply two dimensions. One nondestructive technique relies on the relationship between interplanar atomic spacing and the angle at which an x-ray or neutron beam is diffracted when it passes through the material. The presence of locked-in elastic strains causes a change in the interplanar spacings. It is possible then to apply x-ray or neutron diffraction equations that relate the diffraction angle, wave length of the x-ray beam, and the interplanar spacing to calculate the distortion. The technique is outside the scope of this lesson, but it is discussed in texts cited in the References at the end of this lesson. There are other nondestructive techniques for the measurement of residual stresses: the Barkhausen (magnetic) technique in which residual strain alters magnetic domains in the material, ultrasonic techniques in which a sound wave is passed through the material, holographic techniques in which the laser replaces strain gages, and so forth. There are varying levels of sophistication in these tests and varying sources of error. We will not discuss any of these procedures in detail. SAE Specification J784A describes the procedure for x-ray analysis and The Handbook of Measurement of Residual Stresses (Ref 10) describes other procedures. As noted in the introduction, residual stresses are the result of locked-in elastic strains in a component. These locked-in strains can be created due to nonuniform plastic deformation (mechanical) and by a nonuniform temperature distribution in a component (thermal). In composite materials having different elastic constants, residual strains can be created by the difference in elastic constants. Finally, some electrochemical processes create residual stresses. See Table 1. Mechanical Residual Stresses Nonuniform plastic strain during cold forming is caused by two opposing effects: friction forces at the tool/die interface and a plastic-strain gradient
Table 1
Thermal

Sources of Residual Stresses

Factors That Can Affect Residual Stress


Metallurgical Mechanical Chemical

Fabrication with heat Welding Flame cutting Hot forming Casting Shrink fitting Operations at high temperature Electrical discharge machining

Heat treatment Stress relieving Annealing Hardening Tempering Diffusion treatment Carburizing Carbonitriding Nitrocarburizing Cyaniding Nitriding Decarburizing

Machining, grinding, and polishing Mechanical surface treatments Shot peening Surface rolling Hammer peening Ballizing Cold forming Stretching Drawing Upsetting Bending and straightening Twisting Autofrettage Interference fitting Service overloads Explosive stressing Cyclic stressing Wear, fretting, bruising, gouging, and cracking

Etching Corrosion Chemical machining Surface coating and plating

50 Types of Failure and Stress

through the section. Friction at the tool/die interface causes material at the interface to flow slower than material away from the interface. More heavily deformed regions of the material flow faster than less heavily deformed regions. In rolling, although there is an overall height reduction, the strain in the rolling direction typically varies through the thickness, especially with light reductions or when the rolls are small. This strain gradient is apparent from a visual examination of the front of the stock exiting the rolling mill. The end is not square (Figure 35). Two different situations are possible: heavy friction tends to make the centerline material flow more than material in contact with the die, giving condition (a); and small reductions and/or small rolls lead to condition (b). If the strain gradient in (b) becomes extreme, the residual stresses are relieved by crack formation in the material with the result that the material may contain hidden cracks along the centerline of a swaged bar (Figure 18). If the notch created in (b) becomes extreme, the material can split along the centerline (allligatoring). The length of the material can also vary across the width, which can cause a variation of residual stress in the rolling direction across the width due to bending stresses and crown in the rolls. Similarly, the expansion in the width direction may be nonuniform through the thickness causing a variation in residual stress This is more pronounced in heavy sections than in thin sections. Case (b) shows that the surface residual stress in the longitudinal direction can be compressive. This effect is utilized to introduce desirable residual stresses in material after manufacturing by final skin rolling and burnishing, which introduce plastic strain only in the surface of the material and therefore produce surface compressive residual stresses. Beneficial residual stresses can be introduced into both coil and leaf springs and in torsion rods by plastically deforming the component to a

Figure 35. Development of residual stress in rolled products, showing strain variation in the rolling direction through the thickness. (a) Typical of large rolls, large friction, and heavy reductions. Tensile surface residual stresses in the rolling direction and compressive residual stress along the centerline. (b) Typical of small rolls, low friction, and light reductions. Compressive surface residual stress in the rolling direction and tensile residual stress along the centerline.

Residual Stresses 51

larger load than it is designed to see in service. Prior to service, compression coil springs are preloaded until they close, creating a tensile residual stress in the spring, which then algebraically adds to the applied compressive service stress, decreasing the net stress in the component. Torsion rods are loaded in torsion until essentially the total cross section has yielded. This produces a surface residual shear stress opposite in sign to the initial twist direction and a residual stress in the interior of the rod acting in the same direction as the original twist. Thick-walled cylinders (i.e., gun barrels) are pressurized internally until the outer fibers of the material just deform plastically in a process known as autofrettage. The part is then machined to final shape, causing redistribution of the internal stresses. The net result is a compressive hoop stress at the inner diameter. In service, the net hoop stress is the sum of a positive applied stress and a negative residual stress. Shrink Fits. Consider the loading condition shown in Figure 36 where a collar (hollow cylindrical section) is shrink fitted onto a solid shaft. Assembly would typically consist of manufacturing the collar to have a diameter slightly smaller than the shaft. The collar is then heated, causing it to increase in circumference and diameter, and the hot collar is then slipped over the shaft. When the collar cools, it exerts stresses on the shaft and vice versa. The collar tries to make the shaft diameter smaller, and the shaft tries to increase the diameter of the collar. The hoop or circumferential stress that exists in the system is shown in Figure 36(b). If the collar is too narrow, the tensile stresses can cause the collar to flow plastically. Alternatively, if the shaft is hollow and the wall too thin, the hoop stress can cause the tube to collapse. Shot Peening. Processes such as skin rolling cannot be used to introduce beneficial residual stresses into components of complex geometry. However, another process known as shot peening can be used. Balls of material are impacted on the surface of the component at high velocity, causing plastic deformation in a thin surface layer. If the ball creates an indentation in the surface (decrease in thickness at the surface), the material must elongate in a plane perpendicular to the compression. Thus,

Figure 36. Shrink fitting a collar on a cylindrical section (a) results in residual stresses (b).

52 Types of Failure and Stress

material flow occurs parallel to the surface at the surface, but not below the surface. The result is a surface residual compressive stress. (Remember: the part of the material that was stretched the most ends up in residual compression.) Variables in the peening operation include the size of the shot, the yield strength of the shot relative to that of the material being treated, and the velocity of the shot. It is possible to shot peen too severely, resulting in a degraded condition. Consider the two materials in Figure 33 that have been peened. For material (a), the peened depth is shallow and the residual stress is not as high as in (b). If a bending load is now applied to both materials, the result is as shown. Material (b) peened to a greater depth (and higher residual stress on the surface) now can fail below the surface because of the higher subsurface residual tensile stress. Figures such as this one can always be sketched to determine the effect of a given residual stress pattern in conjunction with an applied loading pattern. It is important to remember that the areas of the tensile residual stress and of the compressive residual stress in the sketch must be equal. Therefore, as a compressive residual stress goes deeper into the material, the magnitude of the subsurface tensile residual stress must increase. Shot peening is routinely used industrially to improve the performance of manufactured components, especially those subjected to cyclic loading and for materials sensitive to stress-corrosion cracking in a given service environment. In many manufacturing operations involving steels, there is some decarburization at the surface during hot-working operations. If the component is loaded in service in bending or torsion, the decarburized region sees the maximum stress. Service performance can be improved by peening, which not only cold works the surface layer to raise its strength but also introduces beneficial compressive residual stress. Beneficial mechanical residual stresses may also be produced into material adjacent to small holes by a process known as ballizing. The process involves pressing a ball through a greased hole that is slightly smaller than the ball diameter. The ball stretches material adjacent to the hole, and, therefore, this region contains compressive residual stresses after processing. Thermal Residual Stresses Just as strain gradients can be created by mechanical processes, they can also be created by thermal processes. Material expands as it is heated and shrinks as it is cooled (assuming no phase transformation in the solid). Therefore, when two different parts of a component are at different temperatures, those two parts would like to occupy different volumes. If they can do this, no residual stress is created. However, if the two sections are bonded together or contiguous, constraint is created and residual stresses develop. When a material is heated to an elevated temperature, thinner sections heat faster than thicker sections, and the outside of a heavy section comes to the operating temperature before the interior of the component. The opposite happens during cooling. The actual magnitude of residual stresses created during cooling depends on whether the material

Residual Stresses 53

undergoes a phase change during cooling. The most important of these phase changes is the transformation of a ductile material at elevated temperature to a brittle material at low temperature, for example, the martensitic transformation. The creation of residual stresses without an accompanying phase transformation is discussed first and then the effect of a martensitic transformation is considered. Problems do develop in limited-ductility materials during heating because the differential strain cannot be relieved by plastic flow, resulting in cracking. Cracking during heating is a problem in the heat treatment of tools and dies. Consider the situation shown in Figure 37 in which a heavy section is quenched to room temperature. The outside layers of the material are rapidly cooled by the quench creating high-yield-strength surface layers surrounding a hot, low-yield-strength core. The initial stress distribution is tension in the outer case and compression in the hot core. However, because the yield strength of the core is low, plastic deformation in the core relieves the residual stresses by plastic deformation. Then, the hot interior begins to cool, creating tensile residual stresses in the core and compressive residual stresses in the surface layer. The general conclusion drawn from this example is the metal that cools last contains residual tensile stresses, providing that there is no phase change to increase the volume of the material. Distortion caused by nonuniform heating or cooling can be examined with Figure 38. Consider a heating rate sufficiently fast that a strong temperature differential exists between the heated and nonheated sides of the plate. The hot side wants to expand because of thermal expansion and causes bowing of the plate as shown because the hot material is now

Figure 37. Longitudinal thermal residual stresses created due to rapid cooling in a material that does not undergo a phase transformation during cooling. The outside is cooled, which causes longitudinal contraction. This region then has a high yield strength. The still hot inside material with a low yield strength is plastically deformed by the cold outside case. Then the inside material cools and contracts. The final result is a state of longitudinal compressive stress in the outside region and longitudinal tensile stress in the core.

54 Types of Failure and Stress

Figure 38. Residual stresses created in a plate rapidly cooled from one longer than the cooler material. At this point, there are compressive stresses in the hot material and tensile stresses in the cool material. However, there are also compressive stresses on the outside of the nonheated side because bending occurred. The bending causes the material near the bottom to be loaded in elastic compression because it is on the opposite side of the neutral axis. The compressive stress on the top is relieved because of the low yield strength of the hot material. Now, the hot material cools and would like to shrink, but it is constrained by the cold material below it. The result is that the plate becomes bowed to relieve partially the elastic stresses. The bowing is now in the opposite direction, so that tensile stresses must exist on the bottom. The final residual stress pattern is tension near the two outer surfaces and compression along the centerline. Thermal Fatigue When strain gradients due to nonuniform heating and cooling occur repetitively in the material, cracks can initiate and grow. Each strain cycle uses up some of the total ductility of the material, and failure occurs when the total strain reaches the fracture strain. This process is known as thermal fatigue. Thermal fatigue is a problem in limited-ductility materials that are thermally cycled, but it can occur over a long period of time in even highly ductile materials such as hot-rolled AISI 1040. If the thermal conductivity of a material is not high and/or if the thermal gradients are large, this becomes a dominant failure mechanism. Thermal fatigue is often characterized by a very large number of small cracks in the fatigued area (Figure 39), although, in other instances, only one or two large cracks may form. Unfortunately, this cracking can occur in the interior of the component, and, therefore, it is not visible. Cooling with Phase Transformations Ferrous materials cooled from above the A3 temperature (upper critical temperature) can transform (and in this order) to: primary ferrite, pearlite,

Residual Stresses 55

Figure 39. Thermal fatigue cracking in tubing. Note the large number of fine cracks on the surface of the tube. Source: Ref 11 bainite, and martensite. The microstructure may also contain untransformed (retained) austenite in the material after quenching, especially for high alloy content and quenching to room temperature. Ferrite has a bodycentered-cubic (bcc) space lattice with a packing factor of 0.68, whereas austenite is face-centered cubic (fcc) with a packing factor of 0.74. Therefore, there is an expansion in steels when fcc austenite transforms to bcc ferrite during cooling, or a contraction when they are heated. Martensite is body-centered tetragonal (bct), and a volume expansion of about 3 to 4% occurs when austenite transforms to martensite. Transformation of austenite to pearlite or bainite also involves an expansion. When austenite transforms to ferrite (and to some extent, to pearlite), the transformation temperature is high enough that the new constituent has a low yield strength so that material can plastically deform to relieve any transformation stresses. That is not the case with martensite (and, to some extent, with bainite). Quench cracking can result if the hardenability of the material is high, including a large grain size. During quenching, the outside of the material cools first, and martensite is formed. This expansion to martensite partially relieves the residual compressive residual stresses in this region that were formed by the thermal gradient. Now, the hot interior begins to cool. If the hardenability is high enough, material just under or inside the previously formed martensite will now also transform to martensite. This expansion due to the second layer of martensite superimposes a tensile stress in the outer layer of martensite. These tensile stresses cannot be relieved by plastic flow in the brittle martensite. Therefore, if the stress reaches the fracture stress of the martensite, quench cracking results. If there is no quench cracking, the residual stress pattern can still be opposite to that obtained when only thermal stresses are present. (For mediumcarbon, low-alloy steels, the residual stresses are compressive on the surface.) Anything that increases the possibility of the formation of deep subsurface martensite increases the possibility of quench cracking and/or surface residual tensile stresses. Quench severity (H value), high hardenability (therefore high carbon content, high alloy content, large prior austenitic grain size), and geometric constraint such as sharp corners are factors that lead to quench cracking. As just noted, a large austenite grain size also promotes quench cracking. If the carbon content is not high and the quench is not too severe, the material may self temper during cooling, reducing the brittleness of the fresh martensite.

56 Types of Failure and Stress

Chemical Residual Stresses Chemical residual stresses are created in plating operations (Table 1), and pre-existing residual stresses may be relaxed during chemical polishing and etching operations, causing distortion/warpage. Electroplating operations can cause other serious material imperfections. The electroplated material contains residual tensile stresses after plating. If this material is not ductile (as in chrome plating), the residual stresses are relaxed by the formation of cracks in the plating. These cracks penetrate to the interface between the plating and base material and act as local stress concentrators and crack initiation sites, especially in cyclic loading. Additionally, plating operations can introduce hydrogen into the material that can subsequently result in hydrogen damage. Specifications often require, especially for high-strength base alloys, that the base material be shot peened before the plating operation so that the tensile stresses produced in the plating operation are minimized. These issues are considered again in Lessons 3 and 4.

Residual Stresses in Surface-Treating Processes

It has already been indicated that cold-working operations can introduce residual stresses in a component. Other common surface-treating processes for ferrous alloys include carburizing, nitriding, and induction hardening. The first two processes change surface chemistry by introducing carbon or nitrogen into the surface of the material. Induction hardening does not change the alloy composition, but the material is processed to have a high-hardness, martensitic case of the same carbon content as the pearlitic core. Because carburizing is performed at a temperature that exceeds the upper critical temperature for the case, cooling from the carburizing temperature involves not only thermal residual stresses, but also phase transformation stresses if the material is quenched from austenite. The phase transformation of interest is usually the transformation of austenite to martensite, rather than to pearlite. The resulting residual stress pattern is shown in Figure 40(a). The starting microstructure for nitriding operations is a tempered martensitic structure, and nitrogen is diffused into the surface of the material at a temperature of about 950 F (510 C), that is, below the upper and lower critical temperatures. Nitrogen introduced into the material forms alloy nitrides that increase the volume of the case. Because the process is carried out below the critical temperatures, there are no phase transformation stresses involving austenite. As the nitriding temperature is also lower than the carburizing temperature, the total thermal gradient is lower, with smaller thermal residual stresses being created. As a result, the formation of alloy nitrides is the primary source of the residual stress pattern that develops (Figure 40a). In the induction-hardening process (or the similar flame-hardening process), only the surface of the material is heated. The process is controlled so that austenite is not created in the interior of the component. At temperature, the outside layer is at a temperature greater than the upper critical temperature. There is an adjacent layer of material heated to elevated temperature but below the upper critical temperature, and the core

Residual Stresses 57
+ FHAIIE 6A IE

+=H>KHE A@ ` 6A FAHA@ =HJA IEJA ?=IA B DECDAH  + = @ FA=H EJE? ? HA EJHE@A@ `6A FAHA@ =HJA IEJA EJHE@AI E ?=IA JA FAHA@ =HJA IEJA ? HA
=

FHAIIE

6A IE  !

 ` 6 ) ! =JAHE= GKA ?DAI J =HJA IEJA ` 6 )! ! ` 7 DA=JA@ =JAHE= FA=H EJE?


>

Figure 40. Axial residual stresses in surface-treated components. (a) induction or flame hardening. (b) Shallow depth carburizing or nitriding. Shot peening residual stresses are the same as those shown in (b). Source: Ref 12.

(interior) is not heated substantially. The component is then cooled rapidly, so that the surface region is quenched to martensite. The resulting residual stress pattern is shown in Figure 40(b). The previous discussion has indicated how thermal and mechanical strain gradients create residual stress. The following list indicates the variety of manufacturing processes that can create residual stresses and Table 1 categorizes the sources of residual stresses. Residual stresses can be created by:

Machining and grinding operations. Differential cooling rates created in heavy sections cooling from an elevated temperature. Differential cooling rates created by cooling one side of component faster than the other side. Phase transformations that create volume changes in materials cooling from an elevated temperature. The volume change may be due to the creation of a new phase in the microstructure, or due to a volume change in the matrix due to a phase change such as the transformation from austenite to martensite in the heat treatment of

58 Types of Failure and Stress

steels; it also includes the surface-treating processes of nitriding, carburizing, and induction hardening.

Some types of hydrogen embrittlement in which hydrogen gas precipitates from solid solution. Electrical discharge machining, which locally vaporizes material causing both thermal stresses and phase-transformation stresses. Welding and brazing operations. Welding of ferrous parts is of particular concern because residual stresses are created both by local heating as well as the transformation of austenite to martensite with a large prior austenite grain size. Electroplating operations. Electroplating produces tensile residual stresses in the member, and material specifications may require material of high hardness to be shot peened prior to the plating operation to mitigate the residual stresses created by plating.

Residual stresses may be locked into the material, or they may be released. Quench cracking is nothing more than the release of residual stresses via fracture in a material having little inherent ductility. In more ductile materials, release of residual stresses may cause warpage or distortion, especially in components in which one dimension is significantly smaller than the other two dimensions. Large-surface-area thin sheets are especially prone to warpage.

Summary

Considering those failures that occur due to distortion or by fracture, issues of importance include:

The type of loading to which the component is subjected: axial (tension or compression), torsion, and bending. Also of importance is whether the load is constant with time, increases with time, or varies periodically with time. The presence of residual stresses. The internal state of stress created by the loads and the part geometry. These stresses may be altered locally by the presence of imperfections and/or the presence of residual stresses. The inherent strength, ductility, and toughness of the material, and the possible variation in those properties with location in the material. The operating environment to which the component is subjected. The service temperature may be high, low, or moderate. Temperatures greater than about 40% of the melting point are considered high and require consideration of creep behavior (discussed in Lessons 3 and 7). Low temperatures are of importance because of the rapid loss in toughness that occurs in materials having a bcc space lattice (and therefore includes the steels), as well as in some materials having a hexagonal close-packed space lattice (e.g., titanium and its alloys and zirconium and its alloys).

References

1. Adapted in part from a presentation by H. Nelson, ASM Fall Meeting (Indianapolis, IN), 1998

References 59

2. D.J. Wulpi, Understanding How Components Fail, 2nd ed., ASM International, 1999 3. N.H. Polakowski and E.J. Ripling, Strength and Structure of Engineering Materials, Prentice-Hall, 1966 4. W.D. Pilkey, Petersons Stress Concentration Factors, 2nd ed., Wiley-Interscience, 1997 5. R.C. Juvinall and K.M. Marshek, Fundamentals of Machine Component Design, 2nd ed., John Wiley, 1991 6. A.G. Atkins and Y.-W. Mai, Elastic and Plastic Fracture, Halsted Press, 1985 7. S.L. Semiatin and J.J. Jonas, Formability and Workability of Metals, American Society for Metals, 1984, p 138 8. R.L. Mott, Applied Strength of Materials, 2nd ed., Prentice-Hall, 1990, p 448 9. A. Blake, Practical Stress Analysis in Engineering Design, Marcel-Dekker, 1982 10. Handbook of Measurement of Residual Stresses, Society for Experimental Mechanics, Fairmont Press, 1996 11. W.T. Becker, Thermal Fatigue Failure in a Vaporizer, Handbook of Case Histories in Failure Analysis, Vol 2, K.A. Esaklul, Ed., ASM International, 1992, p 112 12. J.O. Almen and P.H. Black, Residual Stresses and Fatigue in Metals, McGraw-Hill, 1963

A. Blake, Practical Stress Analysis in Engineering Design, Marcel Dekker, 1982 C. Brooks, Heat Treatment of Ferrous Alloys, McGraw-Hill, 1979 J.A. Collins, Failure of Materials in Mechanical Design, 2nd ed., Wiley Interscience, 1993 Failure Analysis and Prevention, Vol 11, 9th ed., Metals Handbook, ASM International, 1986 Fractography, Vol 12, 9th ed., Metals Handbook, ASM International, 1987 Handbook of Measurement of Residual Stresses, Society for Experimental Mechanics, Fairmont Press, 1996 C. Lipson and R.C. Juvinall, Handbook of Stress and Strength, Macmillan Co., 1963 Residual Stress for Designers and Metallurgists, L.J. Vande Walle, Ed., ASM International, 1981 D.J. Wulpi, Understanding How Components Fail, 2nd ed., ASM International, 1999

Selected References

Metric Conversion Factors


To convert from To Multiply by

in. in. mil in. 2 in. 3 in. ft 2 ft 3 ft oz lb Btu Btu/lb F Btu/ft h F in./in. F psi psi ksi ksi 1/2 ksi in. ksi in. ozf lbf lbf ft lbf 2 lbf/in. 3 lbf/in. 3 lb/ft 3 lb/in. 3 lb/in. gal (U.S. liquid) gal (U.S. liquid) lb/gal ft/gal F F C C K
12 10 9 10 6 10 3 10 2

mm m m m 2 m 3 m m 2 m 3 m g kg J J/kg K W/m K m/m K Pa kPa kPa MPa 1/2 MPa m MPa m gf kgf N N m (or J) 2 kgf/cm 3 kgf/m 3 kg/m 3 g/cm 3 kg/m L 3 m g/L ml/L C K F K C
Multiple and submultiple units
1

25.4 3 25.4 10 25.4 25.4 4 6.45 10 5 1.64 10 1 3.048 10 2 9.29 10 2 2.831 10 2.834 10 1 4.536 10 3 1.054 10 3 4.18 10 1.730 1.8 3 6.895 10 6.895 3 6.895 10 6.895 1.099 1.099 28.4 1 4.536 10 4.448 1.356 14.223 4 2.768 10 16.019 2.768 10 4 2.768 10 3.785 3 3.785 10 119.826 748 (F 32)/1.8 (F 459.67)/1.8 (C 1.8) 32 C 273.15 K 273.15 10 2 10 3 10 6 10 9 10 12 10 deci centi milli micro nano pico

10 10

tera giga mega kilo hecto deka


Abbreviations

J kgf L

joule kilogram force liter

m mm N

meter millimeter Newton

Pa K W

pascal kelvin watt

ASM International Materials Park, OH 44073-0002 www.asminternational.org

ASM International Materials Park, OH 44073-0002 www.asminternational.org

S-ar putea să vă placă și