You are on page 1of 206
THE UNIVERSITY OF CALGARY Mechanistic Studies on Stress Corrosion Cracking of Pipeline Steels in Near-neutral pH Environments by Biao Gu A DISSERTATION SUBMITTED TO THE FACULTY OF GRADUATE STUDIES IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY DEPARTMENT OF MECHANICAL & MANUFACTURING ENGINEERING CALGARY, ALBERTA. DECEMBER, 1999 © Biao Gu 1999 ‘National Library of Canada Bibliotheque nationale du Canada Belographe Services Sertices bobographiques Sanwa On AOR Seine ON IA ON Sonata Sirus ecm vee The author has granted a non- L’auteur a accordé une licence non exclusive licence allowing the exclusive permettant ala National Library of Canada to Bibliotheque nationale du Canada de reproduce, loan, distribute or sell reproduire, préter, distribuer ou copies of this thesis in microform, vendre des copies de cette these sous paper or electronic formats. la forme de microfiche/film, de reproduction sur papier ou sur format électronique. The author retains ownership of the _L’auteur conserve la propriété du copyright in this thesis. Neither the droit d’auteur qui protége cette these. thesis nor substantial extracts from it Ni la these ni des extraits substantiels may be printed or otherwise de celle-ci ne doivent étre imprimés reproduced without the author’s ou autrement reproduits sans son permission. autorisation. 0-612-49499-3 Canadi ABSTRACT Experimental and theoretical approaches towards the study on mechanism of stress corrosion cracking (SCC) of pipeline steels in near-neutral pH environment have been performed. The subject of the investigations included SCC susceptibility at various testing conditions, hydrogen distribution around SCC crack tip and the role of hydrogen in cracking process. SCC tests were mainly conducted using slow strain rate testing (SSRT). The hydrogen distribution around a stress corrosion crack tip was measured using secondary ion mass spectrometry (SIMS), and modelled using an elastic-plastic analysis. A thermodynamic model was proposed and used to calculate the effect that the presence of hydrogen and stress has on the SCC growth rate. SSRTs showed that transgranular stress corrosion cracking (TGSCC) of pipeline steels could occur in a dilute bicarbonate solution with a near-neutral pH value. SCC susceptibility increased as the applied electrochemical potential, the pH value of solution and applied strain rates during tension testing decreased. Hydrogen precharging or addition of carbon dioxide (CO;) facilitated the process of SCC, suggesting that dissolution and ingress of hydrogen are both involved in the cracking process. SIMS measurement indicated that low pH values and applying cathodic potential facilitate the generation, evolution and enrichment of hydrogen around the SCC crack tip. Hydrogen plays an important role in the SCC of pipeline steels by promoting anodic dissolution and SCC susceptibility. ‘Thermodynamic analysis showed that hydrogen interacted with the stress field and changed the internal energy and entropy of the steel. These changes could result in an increase of the anodic dissolution rate of the steel and enhancing the SCC growth rate. The hydrogen-facilitated SCC growth rate obtained from the proposed model was in agreement with SSRT measurements. ‘The mechanism of SCC for pipeline steels in near-neutral pH solutions may be that at near free corrosion potential (Ecar), hydrogen was evolved from environment and accumulated in a highly stressed region of the steel, which enhanced local dissolution or pitting process. The combination of preferential dissolution with stresses induced SCC cracks initiation. Once crack initiated it propagated by hydrogen-facilitated dissolution. At cathodic potentials, when hydrogen concentration reached a critical value, the cracking process is controlled by hydrogen-induced cracking (HIC), ACKNOWLEDGEMENT ‘The author wishes to express sincere gratitude to his supervisor, Dr. X. Mao, for his guidance, support and encouragement throughout this project. Sincere thanks to the technical staff of the department of Mechanical and Manufacturing Engineering, especially A. Moehrle and B. Ferguson for their technical support. ‘The author acknowledges the assistance of Dr. J.L. Luo and Dr. Qiao at the University of Alberta, R. Sutherby of Nova Corp. of Alberta and TransCanada Pipeline and M. J. Wilmot of Nova Research & Technology Corporation during the work. IPSCO Canada, Nova Corp. of Alberta and TransCanada Pipeline are also gratefully acknowledged for supplying the steel used to machine experimental specimens. Finally, the author acknowledges the financial support of the Department of Mechanical Engineering and the Natural Sciences and Engineering Research Council DEDICATION This work is dedicated to my wife, Pili, and daughter, Frances, for their incessant support, and to my parents as a testimony of gratitude. TABLE OF CONTENTS APPROVAL PAGE... ABSTRACT... ACKNOWLEDGEMENT... DEDICATION... TABLE OF CONTENTS LIST OF TABLES ... LIST OF FIGURES.. CHAPTER 1 CHAPTER 2 21 22 2.2.1 Historical Sketch. 2.2.2 Characteristics of SCC... 2.2.3 Sequence of Events in SCC. 2.2.4 Metallurgical Effects... 2.2.5 Electrochemical Effects.. 2.3. The Mechanisms of SCC. 2.3.1 SCC Mechanisms on Anodic Dissolution. 2.3.2 Mechanisms on Hydrogen Induced Cracking .. vil CHAPETER 3 24 EXPERIMENTAL METHODOLOGY. 3.1 32 33 34 2.3.3 General Comment... ‘The SCC of Pipeline Steels. 2.4.1 Field Investigation. 2.4.2 Controlling Parameters in Both Types of Pipeline SCC.46 2.4.3 Mechanisms for SCC in Pipelines. . Preparation of Specimen.. Testing Solutions. Testing Apparatus. 3.3.1 ATM Model TTS-25KN Testing Machine. 3.3.2 SCC Growth Testing System... 3.3.3 Potentiostats.. Experimental Procedure... 3.4.1 Slow Strain Rate Tensile Test... 3.4.2 Electrochemical Polarization Test... 3.4.3 Hydrogen Measurement. 3.4.4 SCC Growth under Low Frequency Cyclic Loading.....82 3.4.5 Fractographic and Microstructure Observation... CHAPTER 4 EXPERIMENTAL RESURLTS AND DISCUSSION - I in Near-neutral pH Environment. Stress Corrosion Cracking Behavior of X-80 and X-52 Pipeline Steels vill CHAPTER 5 41 42 4.3 Discussion... Basic Material Properties. 4.1.1 Mechanical Properties... 4.1.2 Microstructure... SCC Behavior in Near-Neutral pH Environment. 4.2.1 Effects of Electrochemical Potential. 4.2.2 Effect of CO: Additions and pH Value. 4.2.3 Effect of Chloride Ion Concentration. 4.2.4 Effect of Strain Rates... 4.2.5 SCC Growth Rate. EXPERIMENTAL RESULTS AND DISCUSSION - I The Role of Hydrogen in Near-Neutral pH SCC... SA 52 53 54 Hydrogen Distribution at Stress Corrosion Crack Tip. 5.1.1 The Hydrogen Profiles under Different Kt .. 5.1.2 Effects of pH Value and Potential on Hydrogen Distribution 5.1.3 Hydrogen Accumulation at the Inclusions... Effect of Hydrogen on Anodic Dissolution... Effect of Hydrogen on SCC Behaviour of Pipeline Steels. 5.3.1 Results from SSRT on the Effect of Hydrogen. 5.3.2 SEM Observation for Hydrogen Enhancing SCC... Discussion. CHAPTER 6 CHAPTER 7 CHAPTER 8 CHAPTER 9 REFERENCES. THEORETIC MODELING ON HYDROGEN DISTRIBUTION AROUND SCC CRACK TIP... 149 THEORETIC MODELING ON HYDROGEN-FACILITATED ANODIC DISSOLUTION SCC GROWTH.. 161 7.1 Thermodynamic Analysis 163 7.1.1 Corrosion Reaction and Dissolution Current at the SCC Crack Tip 163 7.12. Free energy Change Caused by Interaction between Hydrogen and Stress. 165 7.2 Calculation for SCC Crack Growth Enhanced by Hydrogen 169 and Stress. MECHANISM OF NEAR-NEUTRAL pH SCC OF PIPELINE 175 STEELS. CONCLUSION AND RECOMMENDATION... 9.1 Conclusions... 9.2 Recommendations... Table 2.1 Table 3.1 Table 3.2 Table 3.3 Table 3.4 Table 3.5 Table 4.1 Table 4.2 Table 4.3 Table 5.1 TABLE LIST Characteristics of high pH and near-neutral pH SCC in pipelines... Chemical Compositions of X-80 and X-52 Pipeline Steels (wt%) Mechanical properties of X-80 and X-52 Pipeline Steels. Chemical Composition of NS-4 Solution. Chemical Composition of Nova Solution. Stress intensity factor for each crack in curved specimen. The basic mechanical properties of X-80 and X-52 pipeline Steels... SSRT of X-80 Steel at 6.34x10"/sec in NS-4 solutions. Crack growth rate under low frequency cyclic loading........ ALS Bulk hydrogen concentration in X-80 steel at different testing conditions...129 Fig. 2-1 Fig. 2-2 Fig. 2-3 Fig. 2-4 Fig. 3-1 Fig. 3-2 Fig. 3-3 Fig. 3-4 Fig. 3-5 Fig. 3-6 Fig. 3-7 Fig. 4-1 Fig. 4-2 Fig. 4-3 Fig. 4-4 Fig. 4-5 FIGURE LIST Simultaneous tensile stress, susceptible metallurgical condition, and critical corrosive solution required for stress corrosion cracking... Sequence of events (left to right) in alloys under stress in a corrosive environment... 12 Schematic anodic polarization curve showing zones of susceptibility to 16 stress corrosion cracking. Four-stage model for lifetime of pipeline SCC... 68 Configuration of Slow Strain Rate Tensile specimen and their orientation respect to the pipeline. Configuration of Compact Tensile Specimens... Configuration of SCC Growth Measurement Specimens... Configuration for SCC growth test loading system... Schematic diagram of the testing set-up for slow strain rate tests... ‘Schematic diagram of the polarization cell and wire connection... The schematic diagram of hydrogen measurement under SIMS.. Stress vs strain curve for X-80 steel and X-52 steel pulled in aii Microstructure of X-80 pipeline steel... Microstructure of X-52 pipeline steel... ‘Stress vs strain curves of X-80 steel tested in NS-4 solution and in air...92 Optical microscopy shows a transgranular SCC crack initiated from a pit in X-52 steel at -750mVssce in NS-4 solution bubbled with 5%CO>.......93 Fig. 4-6 Fig. 4-7 Fig. 4-8 Fig. 4-9 Fig. 4-10 Fig. 4-11 Fig. 4-12 Fig. 4-13 Fig. 4-14 ‘SEM fractograph of X-52 steel tested at Exon of -720mVsce in NS-4 solution bubbled with 5%CO> at strain rate of 6.34x10°/sec... 94 SEM micrograph of part of the fracture surface of X-80 SSRT samples tested at Eco of -710mV(SCE) in NS-4 solution bubbled with 5%CO> at strain rate of 7.9x10“/sec... 95 Reduction in area from SSRT on X-80 steel at various potentials in NS-4 solution bubbled with N2 and CO>.. 97 Reduction in area from SSRT of X-52 steel at various potentials in NS-4 solutions with 5% CO2.. 99 ‘SEM photograph for X-80 steel samples tested at (a)-550mVsce, () free corrosion potential, -720mV and (c)-1000mVsce in an NS-4 solution wit 5%COz, --100 The effect of percentage of CO2 added into solution on SCC initiation from SSRT of X-80 steel tested at Ecorr in NS-4 soultion with a strain rate of 2.2x10° /sec... --102 (@) Step-like patterns in the fracture surface of the sample tested at Egor of -700mV in NS-4 solution bubbled with 100% CO) at strain rate of 2.2x10°/sec and (b)fracture surface characteristics near the step-like pattern... 104 Reduction Area of X-80 Steels during SSRT in NS-4 solution at different PH values... 106 ‘SEM pictures for X-80 steel samples. Tested in NS-4 solution with 107 different pH value at free corrosion potential. Fig. 4-15 Fig. 4-16 Fig. 4-17 Fig. 4-18 Fig. 4-19 Fig. 4-20 Fig. 4-21 Fig. 4-22 Fig. 5-1 Fig. 5-2 Fig. 5-3 Fig. 5-4 Effect of CI’ in solutions on reduction in area of X-52 steel at different 110 potentials... Reduction in area from SSRT vs strain rate on X-80 and X-52 steel tested at free corrosion potential in NS-4 solution bubbled with 5%CO2.......111 Multiple cracks initiated in side surface during SCC for X-80samples at Ecomr with 100% CO; in NS-4 solution with different strain rate. A113, Average crack growth rate at different potentials during SSRT..........114 SCC growth of crack3 with K= 22.43 MPam'? during low frequency cyclic loading test (R=0.82, frequency=10 cycles/day)... 16 Discontinuous crack growth with K= 22.43 MPam'? during low frequency cyclic loading test (R=0.82, frequency=10 cycles/day).. 17 Fracture surface of the sample tested at -1000mVsce in an NS-4 solution bubbled with No at strain rate of 6.34x10"/sec ... 121 Individual cracks on the side surface in the sample tested at -1000mVsce in an NS-4 solution bubbled with Nz at strain rate of 6.34x10"/sec......122 Hydrogen concentration near crack tip of X-80 steel with K/=36MPam’? before and after immersion in NS-4 Solution. 125 Effect of stress intensity factor on hydrogen distribution near the crack tip 127 for X-80 steel at free corrosion potential... Effect of pH value of solution on hydrogen distribution near the crack tip for X-80 steel at free corrosion potential... 128 Effect of applied potential on hydrogen distribution near the crack tip for X-80 steel at free corrosion potential. 130 xiv Fig. 5-5 Fig. 5-6 Fig. 5-7 Fig. 5-8 Fig. 5-9 Fig. 5-10 Fig. 5-12 Fig. 6-1 Fig. 6-2 Fig. 6-3 Fig. 6-4 131 Hydrogen accumulation at inclusion of MnS. Effect of precharged hydrogen on polarization Curve of X-52 Steel in bicarbonate solution, (a) in 0.05 M NaHCOs, (b) in 0.005M NaHCO; and (©) in NS-4 solution. 133 Stress vs strain curve for X-80 steel with and without precharged hydrogen in NS-4 solution and in air. 136 Reduction in area of X-80 steel with and without precharged hydrogen vs potentials in NS-4 solution bubbled with pure No... 138 Reduction in area from SSRT vs strain rate on X-80 steel for hydrogen uncharged and precharged samples at free corrosion potential in NS-4 solution bubbled with 5%CO; and tested in air. +139 SEM fractographs for the X-80 steel samples tested with and without hydrogen precharging at free corrosion potential in an NS-4 solution with S5%CO>. 141 ‘SEM photographs for samples pulled in air showing the effects of 142 precharging.. Effect of hydrogen precharing on SCC initiation tested at free corrosion 143 potential in NS-4 solution with 5% CO2, Configuration of compact tensile specimen for hydrogen measurement.151 Hydrogen distribution around crack tip for X-80 steel obtained by HRR 154 15S solution and SIMS measurement. Model for interaction between hydrogen and dislocation... A co-ordinate system with an edge dislocation at its origin. 158 x Fig. 7-1 Fig. 7-2 Fig. 7-3 Fig. 7-4 Fig. 8-1 Fig. 8-2 Fig. 8-3 Model for hydrogen facilitated anodic dissolution type of SCC growth.162 ‘The stress factor Kz on anodic dissolution... 172 ‘The synergistic effect Kye of hydrogen and stress. 173 The relationship between stress corrosion crack growth rate and applied strain rate for hydrogen charged and uncharged X-80 pipeline steel at free potential in NS-4 solution bubbled with 5% CO2. so lT4 Cracks initiated on the side surface of samples for SCC and HIC. (a) tested at free corrosion potential, (b)tested at -1000mVsce. Extent of cracking on the side surface from SSRT samples in NS-4 solution at various potentials and pH values... Schematic illustration of SCC initiation. CHAPTER ONE INTRODUCTION Stress corrosion cracking (SCC) is one of those irritating considerations of the designer who may have to select materials of construction to meet a series of other property requirements that cannot be waived at all or can only be waived within narrow limits. The SCC problem must therefore be considered in the context of the other constraints on design and maintenance, including costs. The designer who must use high strength materials will not be able to select structural alloys which are totally immune to SCC, so that one must understand the meaning of test data characterizing susceptibility to this form of failure. The alloy developer also needs to understand the significance of macroscopic characterization data since theory is inadequate to guide alloy development. SCC is generally defined as the initiation and/or propagation of cracks in the presence of both a tensile stress and a corrosive environment. The occurrence of SCC is commonly recognized by brittle cracking features in normally ductile metals, meaning that the catastrophic capability of SCC is manifested through an effective reduction in fracture toughness. Consequently, SCC is a serious and growing concern in many industries, one of which is the pipeline industry. 2 The occurrence of SCC around world on gas and oil pipelines has been a serious problem. Over twenty years ago, the first incident of SCC initiating on exposed outer surfaces of buried gas transmission pipelines was reported in the United States. More recently, similar events have also arisen in Australia and Canada. The basic characteristics of the Australian pipeline SCC corresponded very closely to the incidents in the United States. However, a very different form of pipeline SCC was found in Canada by both Alberta Gas Transmission Division (AGTD) of NOVA Corporation and TransCanada Pipelines (TCPL). These dissimilarities precipitated the establishment of two categories of SCC in pipeline steels, with the American and Australian version being dubbed “classical SCC” (also referred to high-pH SCC), and the Canadian form becoming “non-classical” SCC. In the past few years in Canada, TCPL, Esso Imperial Oil and Nova Co. experienced over 20 failures due to SCC of pipeline buried underground (Delanty et.al, 1992a). Explosions of pipelines buried underground have occurred due to the leaking of natural gas from stress corrosion crack and threaten public safety. This led the National Energy Board initiate a public safety inquiry into the problem in 1995(National Energy Board, 1996). SCC on Canadian pipelines begins when small cracks develop on the outside surface of the buried pipe. These cracks are initially not visible to the eye and are most commonly found in "colonies", with all of the cracks positioned in the same direction. Over a period of years, these individual cracks may lengthen and deepen and the cracks within a colony may join together to form longer cracks. Since SCC develops slowly, it 3 can exist on pipelines for many years without causing problems. However, if a crack becomes large enough, the pipeline will eventually fail and either leak or rupture. Many Canadian oil and gas pipelines are the longest in North American and are approximately 30-40 years old. Due to the age of the pipelines, pipeline maintenance and replacement costs a lot of money annually. Evaluation and maintenance of aging pipelines relies on the understanding of the mechanism which dominates cracking. The SCC found recently in Canadian pipeline is very unique and mostly occurs in near-neutral pH environments. This is so-called non-classical SCC (also referred to near-neutral pH! SCC). The background in fundamental science of the non-classical SCC under near-neutral pH environment is unknown, Since the low pH values favour hydrogen accumulation in stressed regions of pipelines, hydrogen must be taken into account in non-classical SCC mechanism (Brass, 1996). It is not clear whether anodic dissolution or hydrogen-induced cracking dominates the cracking process. Therefore, it is necessary to carry out a fundamental study on the cracking mechanisms of pipeline steel in near-neutral pH solution, especially a study on the role of hydrogen in cracking process and how it enhances the anodic dissolution type of SCC. In this investigation, we study the mechanism of near-neutral pH SCC of pipeline steels from both experiments and theories. Our work focuses on exploring the role of hydrogen in the cracking process and includes measurement of hydrogen distribution around SCC crack tips and its influence on anodic dissolution and SCC susceptibility. A 4 thermodynamic model of hydrogen-facilitated SCC of anodic type is proposed based on experimental and theoretical analysis. The objectives in this fundamental study investigation are: 1) To study the controlling parameters, such as electrochemical potential, pH value and strain rate, effect on the SCC behavior in near-neutral pH environment so as to identify the critical conditions for low pH SCC it iation in pipelines. 2) To measure SCC growth rate in near-neutral pH environment under slow strain rate and low frequency cyclic loading conditions. 3) To quantitatively measure the hydrogen distribution along a stress corrosion crack tip and investigate the role of hydrogen in the cracking process of non-classical SCC. 4) To develop a micromechanical model to analyse the hydrogen distribution at SCC crack tips and to develop a thermodynamic mode! of hydrogen-facilitated anodic dissolution to provide a fundamental approach for near-neutral pH SCC prediction. Chapter 2 provides a synopsis of the current level of knowledge regarding SCC behavior in general and addresses various mechanisms in SCC, with detailed information given for studies relating directly to pipeline SCC phenomena. This is followed by an s introduction of the experimental techniques and equipment which are used in SCC tests, comprising Chapter 3. Chapter 4 presents the experimental results and preliminary analysis on SCC behavior for pipeline steels in near-neutral pH environment. Chapter 5 is devoted to exploring the role of hydrogen in near-neutral pH SCC processes and providing a detailed discussion of the present findings. Chapter 6 and chapter 7 provide physical models and computation on hydrogen distribution in the SCC crack tips and near-neutral pH SCC growth based on the concept of hydrogen-facilitated anodic dissolution. Chapter 8 addresses a non-classical SCC mechanism which is generated from present experimental and theoretical work. In the final section, Chapter 9, several conclusions are drawn and recommendations are proposed for future research on non-classical SCC study. CHAPTER TWO LITERATURE REVIEW 21 Introduction A general overview of SCC characteristics, fundamentals and various mechanisms is first presented in this chapter. Then current observations and research into both high pH SCC and near-neutral pH SCC of pipeline steels are critically reviewed. This information will provide the conceptual background for the future theoretical modeling and experimentation in non-classical SCC behavior. 2.2 The fundamental Aspects of General SCC Phenomena SCC is defined here as a type of material failure which occurs when certain materials are subjected to tensile stress in a corrosive environment. It requires both an acting tensile stress and the presence of a specific corroding medium. Removal of either will prevent the initiation of cracking or stop the growth of cracks that have already propagated. 2.2.1 Historical Sketch SCC first became a widespread problem with the introduction of the cold drawn. brass cartridge case during the last half of the 19th century. Toward the end of the century it appeared in the brass (but not in the unalloyed copper) condenser tubing in then the 7 rapidly growing electric power generation industry. During this era the problem became sufficiently important to acquire its own name, “season cracking.” By the close of the 19th century the role of residual stresses in causing SCC in brass, was so widely known that stress relieving heat treatments for cold formed products were developed as mitigative measures, and acidified mercury nitrate, which will cause mercury cracking. It was during the late 19th century that ammonia was found to play a causative role in the SCC of brass. This finding contributed to the development of the rule that there is a specificity of environment which will cause SCC in a given alloy. This specificity is usually cited as a prime characteristic of SCC, but the growing number of known exceptions makes the specificity rule of questionable merit. Regardless of the question of specificity, it became obvious that in most cases the responsible species need not be present either in large quantities or in high concentration, at least not in high concentration in the bulk environment. By the end of the 19th century “caustic cracking” of riveted boiler stee! could be cited as another example of SCC and as another indication that the problem might occur in a number of alloy systems. The necessary conditions in caustic cracking are local high concentration of free alkali and elevated temperature. By this point the pattem of failures and mitigative measures had become discernible: there are three elements of the phenomenon, mechanics, metallurgy, and 8 chemistry of the environment, as shown in Fig.2-1. A given problem can be solved only by changing one or more of these three elements, and valid research in the laboratory requires adequate attention to all three. 79 Corrosive environment Fig. 2-1 Simultaneous tensile stress, susceptible metallurgical condition, and a critical corrosive solution are required for stress corrosion cracking Early in the 20th century SCC attributable to atmospheric moisture was seen in aluminum alloys. Also during this period SCC was observed in martensitic steel, but the problem did not become widely recognized for SCC until the era of the aerospace programs. Also early in the 20th century the cracking of mild steel due to nitrates became of practical importance in the chemical industry. From this experience we have a clear 9 statement of an important characteristic of SCC, namely, that often it occurs when the alloy is almost inert to the environment which does the cracking. During the 1930s when stainless steels came to be used extensively in the paper, chemical, and petroleum industries, SCC was observed in this class of alloys, particularly in chloride or caustic solutions at elevated temperature. Also during the 1930s, magnesium alloys for military aircraft were found to be susceptible to SCC in moist atmospheres. During the 1950s the aerospace programs encountered, in addition to SCC of martensitic steels indicated earlier, cracking of titanium alloys in contact with nitric acid, or in contact with salt at high temperatures. During the 1960s, titanium alloys were also observed to be susceptible to SCC in nitrogen tetroxide, in water, and in methanol. Also, a zirconium alloy was found to be susceptible to SCC in iodine bearing environments. These observations supported the growing supposition that SCC is a general phenomenon to be expected in all alloy systems. By 1960, a technique of producing high resolution replicas of fracture surfaces had been developed. Its application to SCC was as inevitable as it was fruitful. It disclosed that on the fracture surfaces there were micron scale details important to the 10 development of theory as well as helpful in critical failure analyses. SCC fracture surface is always macroscopically brittle in appearance; electron fractography demonstrated the gross differences in the fracture topology on the micron scale between it and “brittle fracture” of the same material. The fact that stress corrosion cracks are always macroscopically brittle, even in alloys which are highly ductile in purely mechanical fracture toughness tests, is another characteristic of SCC. By the 1960s, fracture mechanics had matured to the point of being able to quantify the stress field around a stress corrosion crack in a useful way, particularly after the analysis for the bent beam specimen made fracture mechanics not only applicable but also practical for long term tests. In addition to the fracture mechanics analysis, much of the methodology and some of the instrumentation developed in fracture mechanics have been useful in the evolution of SCC testing concepts and procedures. 2. 2 The characteristics of SCC may be summarized as follows: 1. Tensile stress is required. This stress may be supplied by service loads, cold work, mismatch in fit-up, heat treatment, and by the wedging action of corrosion products. 2. Only alloys are susceptible, not pure metals, though there may be a few exceptions to this rule. u 3. Generally only a few chemical species in the environment are effective in causing SCC of a given alloy. 4. The species responsible for SCC in general need not be present either in large quantities or in high concentrations. 5. With some alloy/corrodent combination, such as titanium alloys and crystalline sodium chloride, or austenitic stainless steel and chloride solutions, temperatures substantially above room temperature may be required to activate some process essential to SCC. 6. Analloy is usually almost inert to the environment which causes SCC. 7. Stress corrosion cracking is always macroscopically brittle in appearance, even in alloys which are very tough in purely mechanical fracture tests. 8. There appears to be a threshold stress at least in some systems below which SCC does not occur. 2.2.3 Sequence of Events in SCC In the most general case, if a smooth specimen is placed in a corrosive environment in which it will eventually undergo SCC, the sequence of events is as shown in the top row (Row A) of Fig 2-2. First a corrosion pit forms. There is an important feature of most corrosion pits, the significance of which is not always appreciated, namely, a porous cap of corrosion products which must be removed in order to see the pit itself. This cap impedes exchange between the corrodent within the pit and the bulk corrodent outside the pit, but it permits inward migration of anions such as chloride. In 12 general the pH within corrosion pits also differs from that outside the pit. The function of a corrosion pit in initiating SCC was once widely thought to be purely mechanical, to concentrate stresses. It now appears that the essential function of the pit when it initiates ‘SCC is not primarily mechanical, but rather it is to provide a mechanism for altering the solution chemistry locally to one favorable for SCC. O44 4G (44S LIFE ett aie a ee, Fig.2-2 Sequence of events (left to right) in alloys under stress in a corrosive environment. Materials in Rows A and B pit, and that in Row B is more brittle, materials in Rows C and D do not pit, but the specimen in Row D has a preexisting flaw. Continuing to the right in Row A of Fig. 2-2, representing the passage of time, we see in the third column a stress corrosion crack emanating from the corrosion pit. Assuming the stress is maintained, eventually the stress corrosion crack grows to such a length that the remaining metal ligament snaps in purely mechanical brittle fracture B (fourth column). The combination of stress corrosion crack length and stress required to set off brittle fracture depends upon the fracture toughness of the alloy, and in fact that combination is a quantitative measure of the toughness. Lower strength alloys such as brass, austenitic stainless steels, and the older aluminum alloys are so tough that brittle fracture would not occur in the usual laboratory specimens. The material in Row B is more brittle than that of Row A, and after a short stress time, corrosion cracks grow (third column), and the remaining ligament snaps. Some alloys are so brittle that the corrosion pit itself has been observed to initiate not SCC but brittle fracture, so that although there is a “delayed fracturing,” there is no SCC or other slow crack growth process. The material of Row C does not pit in the corrodent, and no SCC occurs. However, the same material containing a crack-like flaw at the surface, as in Row D, may experience rapid SCC. Titanium alloys in sea water are examples of materials which behave as depicted in Rows C and D. It should be emphasized that neither a preexisting flaw nor a corrosion pit is necessary for initiating SCC if the environment has the correct composition for the alloy. For example, although a titanium alloy may not undergo SCC in salt water except from a preexisting crack or flaw, in methanol SCC initiates readily at a smooth surface of the same alloy. Also the precipitation hardening steel designated “13-8 Mo” will not initiate SCC in neutral salt water under given stress conditions until a corrosion pit is formed. 6 But if the salt water is acidified with HCl, presumably simulating the acidity within corrosion pits in that steel, SCC initiates readily at a smooth surface. The important point of the foregoing discussion is that if one wishes to establish stress corrosion characteristics of a material, he must exclude possible confusion from either pitting or non-pitting behavior on the one hand and brittle fracture on the other. 2.2.4 Metallurgical Effects Pure metals are more resistant to SCC than alloys of the same base metal but are rot immune. For example, cracks have been induced in pure copper in slow strain rate tests, but these conditions are quite severe in comparison to service conditions. Virtually all alloys are susceptible to some degree in the appropriate environments, and susceptibility increases with strength in any given alloy class. However, even alloys of low yield strength are susceptible (e.g. brasses and stainless steels). A complete understanding of SCC requires an explanation of the high resistance of pure metals, as compared to the alloys of the same metals. ‘SCC may be either transgranular or intergranular, but the crack always follows a general macroscopic path that is normal to the tensile component of stress. In transgranular failures, the cracks propagate across the grains usually in specific crystal planes, usually having low indices such as (100), (110), and (210) (Liu, Ret al., 1980). The cracks follow grain boundaries in the intergranular mode. Transgranular failures are 5 less common than the intergranular ones, but both may exist in the same system or even in the same failed part, depending on conditions, Facing halves of recently failed specimens often exactly match one another, whether intergranular or transgranular, indicating that cracking is primarily by mechanical fracture, with little electrochemical dissulution or corrosion during the fracture process. However, anodic dissolution does play a significant role, as discussed in the following Section 2.2.5. The intergranular failure mode suggests some inhomogeneity at the grain boundaries. For example, segregation of sulfur and phosphorous at grain boundaries is the probable cause of intergranular SCC of low alloy steels. And in fact, intergranular stress corrosion cracking may be the result of stress-assisted intergranular corrosion since most alloys showing such failures also show at least weak evidence of intergranular corrosion without stress (Parkins, 1990a). 2.2.5 Electrochemical Effects Electrochemical potential has a critical effect on SCC. Figure 2-3 shows the schematic potentiodynamic anodic polarization curve for a typical active-passive corrosion-resistant alloy, with crosshatched zones where SCC occurs in susceptible alloy- solution combinations. The passive film is an apparent prerequisite for SCC, but the two zones of susceptibility appear at the potential boundaries where the passive film is less stable. In zone 1, SCC and pitting are associated in adjacent or overlapping potential ranges. The common example of zone-1 SCC is austenitic stainless steel in hot MgClz 6 solutions (Staehle, 1977). SCC occurs in a narrow potential range, with pitting present at slightly noble potentials and passivity, with no cracking at slightly active potentials. Although stress corrosion cracks may initiate at pits due to stress intensification, they are not necessarily a prerequisite for SCC, even in zone 1. However, in some instances, potent solutions or oxides that are unstable on the exposed surface can accumulate within pits and initiate cracks. For example, cracking of carbon steels exposed to hot water or nitrates initiates in pits where magnetite can accumulate (Parkins, 1990a). (= POTENTIAL —> (4) 199 CURRENT DENSITY Fig. 2-3, Schematic anodic polarization curve showing zones of susceptibility to stress corrosion cracking (Stachle, 1977) 7 In zone 2, far from the pitting potential range, SCC occurs where the passive film is relatively weak at active potentials barely adequate to form the film. Zone-2 SCC is typified by carbon steel in hot carbonate/bicarbonate solutions (Parkins, 1974). SCC has been observed even in the active region, for example, for carbon steel in strong caustic solutions (Parkins, 1979). However, because anodic currents decrease with time, film formation and growth is probably present even in active potential ranges. It was found that crack growth rates were proportional to anodic dissolution currents at straining electrode surfaces (Parkins, 1979). Agreement between crack growth rates and anodic current densities was generally good. However, some systems, particularly those with fast transgranular, have crack growth rates higher than can be accounted for by simple electrochemical dissolution. The photomicrographs also indicate that topographic features of the brittle crack surfaces have been preserved and thus were not subject to extreme anodic dissolution. Thus, electrochemical anodic dissolution probably initiates mechanical fracture processes (Jones, et al, 1987), accounting for both anodic current proportionality and brittle crack topography. ‘The alterations which occur in the corroding solution within the cavity of a pit while it is growing have already been mentioned. It has been found that the corrodent within stress corrosion cracks in high strength steels, aluminum alloys, and titanium alloys is acid, even though the bulk corrodent was nearly neutral. This acidification may be described in terms of traditional hydrolysis reactions, such as a metal salt reacting with 18 water. Alternatively one can describe the reactions as involving the combining of metal with oxide or hydroxyl ions, leaving an excess of hydrogen ions. These hydrogen ions are balanced by the influx of anions such as chloride. The result is an acid solution which simultaneously may promote the reduction of hydrogen and oppose the repassivation of the metal surface. The acidic nature of the corrodent within stress corrosion cracks in the heavy metals is one reason that effective inhibitors for SCC propagation are virtually nonexistent. It is difficult enough to solve the problem of inhibiting acid chloride solutions. This task is made even more difficult by the exceedingly small amount of corrodent within stress corrosion cracks, the amount being too small to provide a sufficient reservoir for inhibitors. For the same reason, the addition of buffers to the bulk environment makes only a modest change in stress corrosion crack propagation - the buffering capacity of even concentrated buffers is not sufficient to overcome entirely the hydrolytic acidification capacity of the clean metal surfaces of the crack. 2.3 The Review on Mechanisms of SCC Many investigators and commentators have asked the question whether there is a single mechanism which can explain SCC in all the systems in which it has been observed. The prevailing present opinion is that such a common explanation is improbable. Generally there are two main types of mechanisms for SCC in aqueous solution, i.e., anodic dissolution and hydrogen induced cracking. For a given SCC system, 19 whether anodic dissolution or hydrogen induced cracking dominates the cracking process should be analyzed depending on chemical, mechanical, electrical and metallurgical variables. Anodic Dissol Early theories of stress corrosion cracking were concemed with a two-stage mechanism: firstly, the electrochemical reaction produces a stress raising pit and from this, secondly, a crack runs a short distance and the electrochemical reaction then occurs again. Where crack growth is by a dissolution-controlled, then the crack velocity may be expected to be related to the bare metal dissolution rate. Thus, writing Faraday’ second law as a penetration rate, the crack velocity (CV) can be given by Ql) where i, = anodic current density, M = atomic weight of the metal, Z = valence of the solvated species, and F = Farady’s constant. Equation (2.1) represents an upper-bound crack velocity by continuous dissolution. The earlier stages of growth may involve appreciably lower crack velocities than those given by Equation (2.1), and there is evidence of such from laboratory tests. 20 The mechanisms on anodic dissolution that have been advanced fall into the following categories: a)Mechano-electrochemical. This category includes such models as the high strain rate at the tip of the growing crack causing continuous depolarization of the anodic area so that cracking by electrochemical dissolution can continue. Dix (1940) was apparently the first to propose that galvanic cells are set up between continuous intermetallic precipitates and adjacent metal at grain boundaries or through paths within the grains. The ensuing corrosion acts under stress to open up a crack. The applied stress was considered helpful in rupturing surface films, thereby exposing fresh metal at the tip of the crack, allowing the reaction to continue. Support of this view derived from potential measurements showing that intergranular crack sensitive paths in various Al alloys, and also in some other metals, were anodic to the grains. Furthermore, cathodic polarization prevented initiation or the growth of cracks already started. ‘The weakness of the electrochemical theory was uncovered through observations that some metals crack only in specific electrolytes, whereas electrochemical action would in fact be expected in a variety of related chemical media of comparable electrical conductivity. Furthermore there appears to be no reasonable electrochemical explanation 2 why additions of extraneous anions to damaging environments should act as inhibitors for SCC, e.g. chlorides and acetates for carbon steels in boiling nitrate solutions. In other words, simple galvanic effects between tip and walls of the crack alone are not sufficient to explain the cause of fracture. b) Film rupture. This model envisions the cyclic formation and rupture of films of corrosion product. Film rupture or slip-step dissolution model was first mentioned by Dix, and modified by Logan (1952) for various metals including aluminum and magnesium alloys, and by Forty and Humble (1963) for a brass. The common point of view is that a film of brittle surface corrosion product ruptures under stress, allowing progressive exposure of metal underneath to further chemical attack. The film-covered side of the crack may, in some instances, act as cathode. Support of this mechanism was derived from potential ‘measurements or from visible or microscopic examination of metals that had undergone sce. Such a mechanism, however, if applicable at all, is probably not general because (1) SCC in a-brass is observed in absence of visible or suspected films, e.g., in alkaline NH,OH, (2) pure metals on which brittle or cathodic corrosion products form quite readily are immune, and (3) tamished metals in air on which equally brittle films exist do not fail. The film rupture hypothesis is also not plausible for the cracking of metals like 2 stainless steels exposed to boiling concentrated MgC12 solution on which an oxide film would have but short life. Potentiostatic anodic polarization curves for 18-8 in boiling 42% MgCl; solution show no evidence of a passive region (Bamartt and Rooyen, 1961). Furthermore, it was demonstrated that an observed decrease of cracking times for 25% Cr, 20% Ni stainless steels following pre-exposure of unstressed specimens to boiling MgC latest solution is caused not by dissolution of a supposed oxide film, but instead by metallurgical aging of the alloy at the test temperature 154 °C (Uhlig and Sava, 1965). Similarly, alloys of Cu-Au or Ag-Au, on which oxide or other non-ductile surface films do not form, are nevertheless susceptible to stress corrosion cracking in FeCl. Finally, both chloride ions which supposedly break down oxide films, and nitrate ions which supposedly favor oxide films, can act as inhibitors for stress corrosion cracking of mild steel and of austenitic stainless steels respectively. Obviously some factor other than rupture of an oxide or other type brittle film is responsible for SCC in most observed instances of metal fracture. As one of the first proposed stress corrosion cracking (SCC) mechanisms, film rupture theory still receives considerable support. Tensile stress is assumed to produce sufficient strain to rupture the surface film at an emerging slip band, and a crack then grows by anodic dissolution of the unfilmed surface at the rupture site. Most investigators now agree that film rupture is essential to initiate cracking, but considerable controversy exists as to how a stress corrosion crack grows thereafter. For example, transgranular cracking would be expected to grow on the active slip plane, where deformation would 3 cause continual film rupture, contrary to the experimental fact that cracks grow on planes of the type (100) or (110). Parkins(1979) has convincingly demonstrated a correlation between crack growth velocity and anodic current density on straining electrode surfaces . However, a crack growing by electrochemical dissolution at an unfilmed crack tip should leave a relatively smooth, featureless surface, contrary to the crystallographic cleavage (transgranular SCC) and well-defined grain boundary (intergranular SCC) features typically present on the fracture surfaces. Such topographic features on opposing brittle fracture surfaces often match exactly, indicating further that the surfaces were formed by brittle mechanical cleavage-like processes with very little surface dissolution. The proportionality of crack growth rate with anodic dissolution current demonstrated by Parkins does not guarantee that the actual cracking process is electrochemical. Electrochemical (charge transfer) reactions may trigger a brittle mechanical cracking process that accounts for crack growth. Indeed, crack growth in some cases has been observed to proceed discontinuously, in steps (Forty and Humble, 1963), and corresponding periodic crack arrest markings (striations) are frequently present on fracture surfaces (Push, 1985), contrary to the expected smooth increase in crack length expected for electrochemical growth. 4 ©) Stress adsorption. This model envisions the reduction of energy to form anew surface by adsorption of specific species. ‘The mechanism of cracking by this name (Uhlig, 1959) proceeds not by chemical or electrochemical dissolution of metal at the tip of a crack, but instead by weakening of already-strained metal atom bonds through adsorption of the environment or its constituents. The surface energy of the metal is said to be reduced, encouraging the metal to part under tensile stress. This mechanism is related to the simplified Griffith criterion of crack formation which equates strain energy in the metal to the surface energy of the incipient crack area. Since stress corrosion cracks propagate in an apparently brittle manner, albeit at low velocities, the Griffith approach to brittle fracture has often proved attractive in the context of environment-sensitive fracture. Thus, the fracture stress, cc, to cause the spread of an elliptical crack, length 2C, is given by (2.2) where E is Young’s modulus and ys is the surface energy. Clearly, any process that lowers ys will reduce the stress for brittle fracture and ys may be lowered by the absorption of appropriate species at the fracture surface. 2s Various investigators have suggested that metal bonds at the crack tip are weakened by adsorption of critical anions from solution. Uhlig originally proposed that specific aggressive dissolved species adsorb at “mobile defect sites” (Uhlig, 1959), weakening the cohesive bonds between adjacent atoms at the crack tip. This mechanism has some appeal, inasmuch as the described bond weakening is similar to the effect postulated for hydrogen during hydrogen induced cracking (HIC) and for liquid metal atoms during liquid metal embrittlement. Also, the environmental cracking of polymers by specific organic solvents has been thought to occur by such an adsorption mechanism. ‘Adsorption in this mechanism is assumed to be potential dependent, to account for cathodic suppression of stress corrosion cracking below a “critical potential”. Effects of inhibitor species are explained by assuming competitive adsorption between aggressive and inhibiting anions. The effect of a specific dissolved species on SCC of a particular alloy is conveniently explained by assuming specific adsorption of such species to cause bond weakening at the crack tip. ‘An explanation of metallurgical effects on SCC requires the assumption of adsorption at mobile defect sites. The nature and character of such defects have not been specified. Pure metals are thought to be resistant because the defect sites “: « - move into and out of the surface areas at the root of a notch too rapidly for adsorption to succeed” (Uhlig, 1985). No explanation is offered as to why such undefined defects should be more mobile in pure metals than in alloys nor why some adsorbed species inhibit, while others promote, SCC. Furthermore, there is no independent evidence of preferential adsorption 26 is not specific at any at such sites. It has been argued (Bockris, 1977) that adsorption particular potential but continually increases above the zero point of charge where positive and negative charges are neutralized on the surface. An apparent critical potential for cracking may appear simply as the result of anodically controlled crack growth which decreases exponentially with potential. Others (eg Jones, 1987) claim that the model fails to explain measurable, but limited plasticity ahead of an atomically sharp crack tip in a ductile metal and the discontinuous nature of crack propagation. Parkins (1990a) summarized these and other arguments against the stress adsorption. ‘There are various difficulties associated with the “stress adsorption cracking” concept, but the demonstration by many workers that small, although measurable, amounts of plastic deformation are involved in stress corrosion crack propagation creates a particular problem. Where plastic deformation is involved in fracture, Oraowan suggests that the surface energy term in equation (2.2) needs to be modified to take into account the work done in plastic deformation, so that to ys should be added yp, the work for plastic strain. Now yp is greater than ys by a few orders of magnitude (5 kJ/m? as ‘opposed to 5 J/m®), and therefore any reduction in the latter by adsorption will have a negligible effect on the fracture stress. Moreover, in some instances mechanism involving enhanced local plasticity caused by the presence of hydrogen has become apparent in recent years and, for those cases at least, the mechanism of crack growth is hardly 27 consistent with an approach based on Equation (2.2). 4) Atomic surface mobility. Galvele (1987) has suggested that many forms of environmentally induced cracking grow by the capture of surface vacancies at the crack tip and counter-current surface diffusion of atoms away from the crack tip. The mechanism predicts that stress corrosion cracking should be prevalent at temperatures below half the melting point of the metal, Tn, and in the presence of low melting surface ‘compounds, when surface mobility is maximized in preference to bulk diffusion in the metal crystal. The effect of critical environments to cause SCC is explained by low melting compounds-enhanced surface mobility. The model leads to an expression for crack growth on the basis of the surface diffusion of atoms, by the following equation: enp( AEs) 1] @3) where Ds = coefficient of surface self-diffusion of the metal, Z = diffusion path length(~ 10% m), a = atomic diameter, o= maximum stress at the tip of the crack, Ey = hydrogen- vacancy binding energy, a = relative degree of saturation with hydrogen of a vacancy at the tip ofa crack, k= Boltzmann's constant, and T= absolute temperature. 2 Parkins (1990a) points out that there are problems in estimating D, for contaminated metal surfaces, as will almost invariably obtain with SCC in aqueous environments, but Galvele concludes that a Ds value of 10"m’s"' is a safe limit, below which surface diffusion and hence cracking are unlikely. Such a D, will be related to temperature, and the implication is that compounds having melting points below 1200 °C, corresponding to the value of D, quoted above at room temperature, will promote stress corrosion; whereas, compounds with higher melting points are protective. This appears to raise a difficulty in relation to the cracking of iron-based alloys. While predicting that nitrates would promote cracking because the melting point of iron nitrate is very low, it would predict that FeO, is a protective compound, since its melting point is 1597°C. It has already been mentioned that in many instances of SCC in steels, the potential range for cracking is associated with FeO, formation. Galvele surmounts this problem by invoking arguments relating to the breakdown the film, chemically or electrochemically, so that active-passive transitions and potential dependence, for ‘example, are expected to be involved in SCC. In that sense, the surface-mobility model is not essentially different from some considered earlier, with the essential difference being that the experimental determination of relevant electrochemical data is replaced to some extent by the properties of the salts involved. ) Film-induced cleavage. The film-induced cleavage (FIC) mechanism (Sieradzki and Newman, 1985) has been proposed to explain discontinuous transgranular crack growth and high transgranular crack growth rates. FIC postulates the presence of 29 brittle surface films, including dealloyed films on certain alloys and oxides on pure metals and other alloys. A crack growing in the film may propagate further into the underlying metal if it reaches sufficient velocity at the film/metal interface. Anodic dissolution and corrosion are not necessary to propagate cracks but only to form the required brittle surface film. Thus, a relatively low anodic dissolution rate forms a brittle surface film, and a fast-growing crack in the film may penetrate for some distance into the ductile substrate metal before its growth is arrested. The surface film must reform at the crack tip surface before a new burst of brittle crack growth is possible. Thus, a low anodic current at a strained surface can foster rapid, discontinuous, brittle cracking, according to the FIC mechanism. Unfortunately, there is no a priori reason to expect that surface films or layers are sufficiently brittle to support a brittle crack or that such a crack will be sustained after crossing the interface into the underlying ductile matrix. Most passive films are very thin and hydrated. While they can be ruptured, it seems unlikely that they would be sufficiently brittle to support the suppositions of FIC. Dealloyed layers are unlikely to have sufficient adhesion to the substrate or inherent brittleness to sustain cracking. Fritz et al.(1988) observed that a stress corrosion crack could not propagate beyond the dealloyed layer in brass without an externally imposed anodic current. g) Localized surface plasticity (LSP). Metallurgical effects on SCC are not easily explained by models that pertain only to reactions on the surface. It would seem 30 that a valid mechanism must include effects of the corrosive environment on the properties of the underlying metal. Mechanical creep precedes the initiation of SCC (Petit and Desjardins, 197), and is accelerated by anodic currents (Revie and Ublig, 1974). Such anodic currents can originate, of course, from the usual anodic corrosion reactions. Creep, accelerated by anodic current, implies softening by a surface defect structure which attenuates the normal strain hardening that accompanies primary creep. Corrosion induced relief of strain hardening may be one symptom of a critical corrosion effect on metallurgical properties during SCC (Jones, 1985). That is, SCC may result from a defect structure ahead of the crack tip, as proposed in the mechanism, localized surface plasticity ones, 1990). In the LSP mechanism, film rupture supposedly initiates large anodic currents at the rupture site by galvanic coupling of the unfilmed active surface to surrounding noble passive surfaces. The passive film must be in a weakened state to prevent rapid repassivation which would suppress crack growth. These weakened states usually occur at critical potentials near the boundaries of the passive potential region and are probably influenced by the presence of critical anions such as CI”. These galvanic anodic currents produce a softened defect structure locally at the rupture site, which coincides with the crack tip for a growing crack. The softened structure, which would normally deform plastically, can do so to only a limited extent in the microscopic volume ahead of the crack, when constrained by the surrounding material of full strength and hardness. Microstrain within the softened, yet constrained, crack-tip volume produces a triaxial 31 stress state (plane-strain condition), which suppresses plastic slip. Continued strain can only propagate a brittle crack. The condition at the crack tip has been likened to that of soft solder joining the ends of two steel bars (Jones, 1990). Loading perpendicular to the thin solder joint causes brittle failure of the soft solder material due to the triaxial stess state, which prevents plastic deformation of the normally ductile solder. Forty (1963) originated the proposal that vacancies injected into the crack tip miero volume during dealloying cause a defect crack-tip structure responsible for SCC of brass. Efforts to demonstrate embrittlement by vacancy production in foils by quenching or deformation have been unsuccessful (Pugh, 1990). According to the LSP mechanism, this is not unexpected. Mechanically tested foils will nearly always be in a biaxial stress state, which permits easy plastic deformation. Only when the softened material is constrained by the plane-strain condition at the crack tip is it subject to brittle cracking. A vacancy-saturated defect structure is one that may be responsible for SCC. Others may be possible as well, and the LSP mechanism is not restricted to any specific defect structure type. However, it is notable that vacancies have been credited with causing increased creep by anodic reactions (Rieve and Uhlig, 1974) and by cyclic loading (Felter and Sinclair, 1963). The LSP mechanism seems to explain several metallurgically related effects in SCC. Resistance of pure metals is explained by the lower degree of strain hardening in 2 pure metals, as compared to alloys. Thus, extensive deformation and ductility in pure metals are permissible before relief of strain hardening is sufficient to initiate cracks. Highly strain hardenable alloys, even of low yield strength (e.g., stainless steels and brasses), are often highly susceptible. Anodic currents are thought to trigger a brittle mechanical cleavage process, which occurs on prismatic planes of the (100) or (110) type. These planes have the least atomic density and greatest atomic spacing and therefore are most susceptible to cleavage when plasticity is suppressed at the plane-strain crack tip. Discontinuous growth of cracks is explained by a postulated brittle crack burst into the softened crack-tip volume and possibly beyond. The softened crack-tip defect structure must reform at an arrested crack tip before a new brittle burst is possible. 2.3.2 Mechanisms on Hydrogen Induced Cracking Some alloys fail under stress in corrosive conditions because of the entry of hydrogen atoms into the alloy lattice. This phenomenon is called hydrogen-induced cracking (HIC). Numerous investigators have suggested that the brittle nature of SCC can only be accounted for by a mechanism controlled by HIC, which has brittle cleavage-like features very similar to those of SCC. It is generally assumed that hydrogen acts to weaken interatomic bonds in the plane strain region at the crack tip. The result is a singular crack with little branching. 3 The process of HIC involves hydrogen evolution, adsorption, diffusion and embrittlement. Hydrogen may be present from reduction of water or acid by H,0 +e + H+OH" (2.4) or H' +e H, (2.5) in neutral and acidic solutions, respectively. SCC, which relies on an anodic dissolution for at least a part of the propagation Process, can be avoided by cathodic protection. HIC, by contrast, is accelerated by cathodic polarization since this promotes the production of hydrogen. This distinction has been used to discriminate between the two mechanisms. Unfortunately, such a simplified assumption does not always make a clear distinction. In some instances, copious cathodic formation of hydrogen and subsequent measured entry into the metal lattice has stopped the propagation of stress corrosion cracks (Wilde and Kim, 1972). Proponents have suggested that in some way hydrogen may enter the lattice at the crack tip, despite contrary surface potentials. Nevertheless, very deep, growing stress corrosion cracks have been stopped effectively in low-strength ductile alloys by cathodic polarization with accompanying hydrogen evolution. Hence, hydrogen is not generally accepted as the cause of SCC in most low strength ductile alloys. However, HIC has generally been accepted as the controlling mechanism of failure 34 for the less ductile high strength iron, titanium, and some aluminum alloys. Despite extensive study, the mechanisms of hydrogen induced cracking have remained unclear. Of many suggestions, three mechanisms appear to be viable: a) Stress-induced hydride formation and cleavage mechanism ‘A number of metallic systems have demonstrated hydrogen embrittlement resulting from stress-induced formation of hydrides or other relatively brittle phases and the subsequent brittle fracture of these phases. Several types of phases may take part in this failure mechanism, e.g. hydrides, martensitic phases, etc. The basic requirements are that these phases be stabilized by the presence of hydrogen and the crack-tip stress field, and that the phase that forms be brittle. The typical system that exhibits failure by this mechanism forms stable hydrides in the absence of stress, and these hydrides are thermodynamically more stable under the stress and hydrogen-fugacity conditions at the crack tip. In some cases, hydrides can be shown to result from this stress stabilization even when they are not formed in the absence of the stress. Among systems that exhibit hydride embrittlement are the group Vb metals (Nb, V, and Ta), Zr, Ti, and alloys based on these metals. This hydride embrittlement mechanism can be qualitatively described as follows. Under the applied stress, the chemical potential of the solute hydrogen and the hydride are reduced at tensile stress concentrations, such as crack tips. Diffusion of hydrogen to these elastic singularities and precipitation of hydrides at the crack tips then occur. The phase

You might also like