Está en la página 1de 53

KING SAUD UNIVERSITY

COLLEGE OF ENGINEERING

RESEARCH CENTER

Final Research Report No.: 2/426

"EFFECT OF HEAT TREATMENT ON FATIGUE AND TOUGHNESS PROPERTIES OF STEEL"


By

Dr. Magdy El Rayes [PI] Mechanical Engineering Department

Prof. Dr. Said Darwish [CI] Industrial Engineering Department

5/ 1427 6/ 2006

Table of Contents Page Acknowledgement 1. Abstract 2. Introduction 3. Summary 3.1. Research Problem 3.2. Research Significance 3.3. Research Objectives 3.4. Research Methodology 4. Literature Review 4.1. Effect of Heat Treatment on Fatigue Life 4.2. Characteristics of Fatigue Fracture Surface 4.3. Nature of Fatigue Process 5. Experimental Procedures 6. Results and Discussion 6.1. Effect of Heat Treatment on Hardness 6.2. Effect of Heat Treatment on Mechanical Properties 6.3. Effect of Various Heat Treatments on Fatigue Life of 4340 Steel 6.4. Effect of Various Heat Treatments on Fatigue Life of 1040 Steel 6.5. Effect of Various Heat Treatments on Fatigue Life of 4140 Steel 6.6. S-N Curve Equations 6.7. Effect of Quenching and Tempering on Fatigue Life of Different types of Steel 6.8. Characteristics of Fatigue Fractured Surface Subjected to different Heat Treatments 6.8.1. Annealing 6.8.2. Normalizing 6.8.3. Quenching and Tempering 6.9. Impact Toughness 7. Conclusions 8. References 46 46 47 48 49 51 52 42 3 5 7 11 11 11 12 12 14 14 16 20 26 29 29 29 30 33 38 40

ACKNOWLEDGEMENT
The authors would like to express their sincere thanks to the Research Center- College of Engineering- King Saud University for financing and supporting this research. Special thanks are due to all the workers in the Manufacturing Engineering LaboratoryCollege of Engineering- Mechanical Engineering Department for their help in machining all the specimens needed in this work. The engineers in charge of Materials Testing Laboratory and the Material Properties Laboratory- Mechanical Engineering Department are deeply acknowledged for their help and assistance executing the experiments with modern and reliable equipments.

List of Tables Table 1: Chemical composition of materials (wt %) Table 2: Mechanical properties of materials Table 3: Heat Treatment procedures applied to the three steel types. Table 4: Mechanical properties of three types of steel measured after different heat treatments. Table 5: The Constant and exponent n values in Basquin relationship List of Figures Fig. 1: Typical fatigue failures in steel components. Fig. 2: Fracture surfaces at high stress and low stress (schematic) [14] Fig. 3: Striations in an aluminum alloy. [14] Fig. 4: Fatigue failures in the Alexander L Kielland platform. [14] Fig. 5: Slip band with extrusions and intrusions formed on the surface of a grain subjected to cyclic stress. Crack nucleation at intrusion. [14] Fig. 6: Stage I and stage II crack growth at polycrystalline material. [14] Fig. 7: Fatigue crack initiation at an inclusion in a high strength steel alloy. [14] Fig. 8: Fatigue specimen geometry, RA= 1mm Fig. 9: Hardness values for three types of steel with different heat treatments QT: Quenched and Tempered N: Normalized A: Annealed Fig. 10: S-N curve for 4340 steel with different heat treatments Q&T: Quenched and Tempered N: Normalized A: Annealed Fig. 11: Steel 4340, Nital etched, x 1200 Fig. 12: S-N curve for 1040 steel with different heat treatments Fig. 13: Steel 1040, unetched, x 120 a. Quenched and Tempered b. Normalized c. Annealed Fig. 14: Microstructure of steel 1040, Nital etched, x1200 a. Quenched and Tempered b. Normalized c. Annealed Fig. 15: S-N curve for 4140 steel with different heat treatments Q&T: Quenched and Tempered N: Normalized A: Annealed Fig. 16: Steel 4140 with various heat treatments, unetched. x 120 a. Quenched and Tempered b. Normalized c. Annealed Fig. 17: S-N curves for three types of steels in the quenched and tempered condition Fig. 18: Slag Inclusions in the Quenched and Tempered for different steel types, unetched, x120 a. steel 4340 b. steel 4140 c. steel 1040 Fig. 19: Optical micrographs of quenched and tempered steels.x1200 a. steel 4340 b. steel 4140 c. steel 1040 Fig. 20: Macrographs of fractured surfaces of annealed specimens. x 10 Fig. 21: Macrographs of fractured surfaces of normalized specimens. x 10 Fig. 21: Macrographs comparing slow-crack growth zone between normalizing and annealing for the same steel type. x 10 Fig. 22: Macrographs of fractured surfaces of quenched and tempered specimens. x 10 Fig. 23: Impact toughness test results for three types of steels subjected to different heat treatments Q&T= Quenching and Tempering N: Normalizing A: Annealing

1. ABSTRACT
This paper presents the effect of various heat treatments on the fatigue life and impact toughness of AISI/ SAE 1040 [Carbon Steel], 4140 [Chromium Molybdenum steel] and 4340 [Nickel Chromium Molybdenum steel]. These steels were annealed, normalized and quenched [hardened] followed by tempering. The properties of tensile and yield strength, fatigue, hardness, impact toughness and microstructure were evaluated. The experimental results of 1040 steel showed that quenching and tempering treatment increases the tensile strength, hardness and impact toughness when compared to normalizing and annealing respectively. On the other hand, steels 4140 and 4340 had similar behaviour with respect to heat treatment type. Normalizing increased their tensile strength and hardness, whereas, quenching and tempering increases both fatigue and impact properties. Annealing treatment with the three types of steels gave the lowest mechanical properties. It was found that quenching and tempering treatment results a predominantly tempered martensititic structure, whereas, normalizing results either bainitic or ferritic- pearlitic structures depending on the steel type. With all steels investigated, annealing caused softening and coarsening of ferritic- pearlitc structure. The microstructure type and slag inclusions characteristics play an important role in understanding and analysing the mechanism of failure in fatigue testing. Macroscopic investigation of the fractured surface morphology is correlated with fatigue and impact toughness results.

0401 0414 ) - ( 0434 ) - - (. . . 0401 . 0414 0434 . . . ) ( . . . .

2. INTRODUCTION

Fatigue is the response of a material to cyclic loading. It occurs typically at stress levels that are insufficient to cause catastrophic failure under monotonic/static loading. Since between 80 and 90% of engineering failures have been estimated to occur by fatigue, it is a particularly important failure mode in engineering practice. Metal fatigue is caused by repeated cycling of the load. It is a progressive localized damage due to fluctuating stresses and strains on the material. Metal fatigue cracks initiate and propagate in regions where the strain is most severe. The process of fatigue consists of three stages: 1. Initial crack initiation, 2. Progressive crack growth across the part, and 3. Final sudden fracture of the remaining cross section. There are three commonly recognized forms of fatigue: high cycle fatigue (HCF), low cycle fatigue (LCF) and thermal mechanical fatigue (TMF). The principal distinction between HCF and LCF is the region of the stress strain curve where the repetitive application of load (and resultant deformation or strain) is taking place. HCF is characterized by low amplitude high frequency elastic strains. An example would be an airfoil subject to repeated bending. LCF is characterized by high amplitude low frequency plastic strains. If we pull the beams of the tuning fork apart until they are permanently bent we have imparted one half of an LCF cycle. In the case of TMF (present in turbine blades, vanes and other hot section components) large temperature changes result in significant thermal expansion and contraction and therefore significant strain excursions. These strains are reinforced or countered by mechanical strains associated with centrifugal loads as engine speed changes. The combination of these events causes material degradation due to TMF. In engine components where HCF is a concern, turbomachinery designers observe what is referred to as a material's fatigue strength. This is determined by running multiple specimen tests at a number of different stresses. The objective is to identify the highest stress that will produce a fatigue life beyond ten million cycles. This stress is also known as the material's

endurance limit. Gas turbines are designed so that the stresses in engine components do not exceed this value including an additional safety factor. Because of increasing demands imposed on high-performance materials and the disastrous and costly results of component failures, the need for fracture and fatigue research has gained significance and importance. So, too, has the need for designers and engineers to gain an understanding of the fundamental principles of failure investigation and analysis, and the fracture mechanics and fatigue of steels that are widely used in industry. These steels are specifically alloy steels which are mainly characterized by durability which is perhaps the most significant attribute an industry can possess. Industrial examples that require high durability and long fatigue life are: Bolts, Connecting Rods, Hydraulic Clamps, Heavy duty Shafts, Bearings, Gears, Conveyor Rolls, Hydraulic Shafts, Axles, Couplings, etc. The fatigue properties of metals are quite structure-sensitive. However, at the present time there are only a limited number of ways in which the fatigue properties can be improved by metallurgical means. By far the greatest improvements in fatigue performance result from design changes, which reduce stress concentration and from the use of beneficial compressive residual stress, rather than from a change in material. Nevertheless, there are certain metallurgical factors, which must be considered to ensure the best fatigue performance from a particular metal or alloy. Fatigue tests designed to measure the effect of some metallurgical variable, such as special heat treatments, on fatigue performance are usually made with smooth, polished specimens under completely reversed stress conditions. It is usually assumed that any changes in fatigue properties due to metallurgical factors will also occur to about the same extent under more complex fatigue conditions, as with notched specimens under combined stresses.

Fatigue properties are frequently correlated with tensile properties. In general, the fatigue limit of cast and wrought steels is approximately 50 percent of the ultimate tensile strength. The ratio of the fatigue limit (or the fatigue strength at 106 cycles) to the tensile strength is called the fatigue ratio. However, the greater structure sensitivity of fatigue properties, compared with tensile properties, is shown in tests comparing the fatigue limit of plain carbon eutectoid steel heat-treated to coarse pearlite and to spheroidite of the same tensile strength. Even though the steel in the two structural conditions had the same tensile strength, the pearlitic structure resulted in a significantly lower fatigue limit due to the higher notch effects of the carbide lamellae in pearlite. There is good evidence that high fatigue resistance can be achieved by homogenizing slip deformation so that local concentrations of plastic deformation are avoided. This is in agreement with the observation that fatigue strength is directly proportional to the difficulty of dislocation cross slip. Materials with high stacking-fault energy permit dislocations to cross slip easily around obstacles, which promotes slip-band formation and large plastic zones at the tips of cracks. Both of these phenomena promote the initiation and propagation of fatigue cracks. In materials with low stacking-fault energy, cross slip is difficult and dislocations are constrained to move in a more planar fashion. This limits local concentrations of plastic deformation and suppresses fatigue damage. The ability to control fatigue strength by altering stacking-fault energy has practical limitations. A more promising approach to increasing fatigue strength appears to be the control of microstructure through thermomechanical processing to promote homogeneous slip with many small regions of plastic deformation as opposed to a smaller number of regions of extensive slip. In general, quenched and tempered microstructures result in the optimum fatigue properties in heat-treated low-alloy steels. However, at a hardness level above about Rc 40, a bainitic

structure produced by austempering results in better fatigue properties than a quenched and tempered structure with the same hardness. For quenched and tempered steels the fatigue limit increases with decreasing tempering temperature up to a hardness of Rc 45 to Rc 55, depend on the steel. The fatigue properties at high hardness levels are extremely sensitive to surface preparation, residual stresses, and inclusions. The presence of only a trace of decarburization on the surface may drastically reduce the fatigue properties. Only a small amount of nonmartensitic transformation products can cause an appreciable reduction in the fatigue limit. The influence of small amounts of retained austenite on the fatigue properties of quenched and tempered steels has not been well established. The fatigue limits of quenched and tempered low-alloy steels of different chemical composition are about equivalent when the steels are tempered to the same tensile strength. This generalization holds for fatigue properties determined in the longitudinal direction of wrought products. However, tests have shown that the fatigue limit in the transverse direction of steel forcing may be only 60 to 70 percent of the longitudinal fatigue limit. It has been established that practically all the fatigue failures in transverse specimens start at nonmetalic inclusions. Nearly complete elimination of inclusions by vacuum melting produces a considerable increase in the transverse fatigue limit. The low transverse fatigue limit in steels containing inclusions is generally attributed to stress concentration at the inclusions, which can be quite high if an elongated inclusion stringer is oriented transverse to the principal tensile stress.

10

3. SUMMARY
3.1. Research Problem: The fatigue properties of steels are quite structure-sensitive. However, presently there is limited number of ways in which the fatigue properties can be improved by metallurgical means such as heat treatment. In common, most researches done in this field have focused only upon case hardening using either quenching and tempering, carburizing, carbonitriding or laser treatment. This resulted in relatively improving the fatigue life than the untreated steels, but was accompanied by loss in toughness. Furthermore, the fabrication-induced defects such as poor surface roughness, inclusion, micro-voids and stress concentrations exist internally and within the entire materials cross section in which neither case hardening effect nor design change might hinder their negative influence on fatigue life. Finally, the surface roughness of the metal plays a significant role in reducing the fatigue life. 3.2. Research Significance: Fatigue is the response of a material to cyclic loading. It occurs typically at stress levels that are insufficient to cause failure under monotonic/static loading. It is a particularly important failure mode in engineering practice. Fatigue effects often cause failures which remain undetected until a catastrophic fracture occurs, frequently, at a very inopportune time and without warning. These failures drive up warranty costs, and the worst case is the potential for devastating financial losses and even endangering lives. Therefore, the evaluation and study of fatigue behavior can no longer be considered optional since between 80 and 90% of engineering failures have been estimated to occur by fatigue.

11

3.3. Research Objectives: This project aims to study the effect of heat treating the full section by quenching [hardening] followed by tempering, normalizing and annealing on the fatigue properties and impact toughness of three different types steels, namely AISI/ SAE 1040 [Carbon Steel], 4140 [Chromium Molybdenum steel] and 4340 [Nickel Chromium Molybdenum]. These treatments were not studied before in earlier literature. Also, the aim of this work is to gain an understanding of how various heat treatments affect failures that originate in different alloy steels, learn the principles and fundamentals in failure investigations, and the basic steps in analyzing the failure of a part. In addition, understanding the inter-relationship between fatigue life, mechanical properties, micro structure and fracture toughness is another objective in this work. Finally, the correlation between fatigue life and fractured-surface morphology using failure analysis is also studied. 3.4. Research Methodology In order to achieve the above-mentioned objective, the following steps will be done using the selected types of steel in the form of bars. 1. Base material chemical analysis, microscopic and hardness testing in order to ascertain the steel type that shall be used in experiments. 2. Machining of fatigue and impact specimens to standard dimensions and surface finish, and also few additional specimens for microscopic and hardness testing, each of which corresponds to the specified heat treatment. 3. Heat treating the specimens prepared in steps 2 and 3. These heat treatments are namely; full hardening of the section [quenching] followed by tempering, normalizing, and annealing.

12

4. Microscopic and hardness testing across the whole section of the additional heat treated specimens [mentioned above in step 2] in order to assure that the heat treatment has been done as planned and to be compared with the original base metal. 5. Fatigue testing using a rotating-bending fatigue testing machine. The number of specimens corresponding to each type of heat treatment will be around 5. This will help drawing the S-N curve with high sensitivity. The rotating-bending fatigue testing is classified as a high cycle fatigue (HCF) regime testing. This type of testing results from vibratory stress cycles at high frequencies, that can reach thousands of cycles per second. The stresses in this type of testing are predominantly elastic, and stress levels are below the yield strength of the work piece material. Test results shall be plotted as S-N and toughness curves for each type of heat treatment. In addition, the fractured surfaces of fatigue specimens shall be studied under high magnification in order to analyze the failure mechanism. 6. Impact toughness testing in which the results will be presented as bar-charts. 7. Evaluating the proposed heat treatments from the fatigue and impact toughness properties points of view and studying the relationship between them.

13

4. LITERATURE REVIEW
4.1. Effect of Heat Treatment on Fatigue Life The influence of heat treatment on fatigue life for various types of steels has appeared in numerous publications. It has been found that the majority of the work has focused upon modifying the surface properties either by heat treatment such as carburizing, nitrocarburizing, laser beam treatments, shot peening surface-treated steels or by depositing a thin layer to enhance either the corrosion-fatigue or contact-fatigue lives. The influence of nickel-phosphorus deposits on the corrosion-fatigue properties of quenched and tempered AISI 1045 steel was studied [1]. It was reported that there is no substantial differences found between fatigue life of the coated and uncoated specimens. Post heat treating the deposited layer in further work [2] reduced significantly the fatigue life of the same base material. On the other hand, nitro-carburized M-50 steel was tested using rolling contact-fatigue [3]. Nitro-carburizing techniques produced a thin compound layer of about 1mm within nitrogen-rich diffusion zone. This increased the rolling contact-fatigue life by a factor of 6 when compared to the untreated material. The effect of inclusions on fatigue properties of 3 different grain size 42Cr Mo V Nb steel was studied [4]. No definite endurance limit was obtained in these tests due to high scattering of results. However, crack initiation sites were related closely to the fatigue life and most fractures originated [initiated] at the inclusions. Although slow cooling rates result in a better strength and ductility of dual phase [ferrite-pearlite] steels, it does not provide any remarkable benefit in regard to low cycle fatigue properties of C-1.2 Mn- 0.6 Si or similar types of dual phase steels [5]. Low power laser beam treatment has been also used to a certain extent as a localized hardening heat treatment method. Martensite formation due to high cooling rate was the

14

result of this treatment which enhanced the fatigue properties and hardness. This is due to the transformation residual compressive stresses at the surface [6]. In addition, laser beam cladding has not only improved the fatigue strength of an alloy steel but also forms a heat, corrosion and wear resisting layers [7]. But this was accompanied with large scatter in fatigue life results and crack growth rates due to some defects such as voids found in the cladded layer. However, fewer defects significantly led to more strengthening. One of the important products of steels heat treatment is the retained austenite. [8]- [9] It has been reported that higher levels of retained austenite of SAE 8620 steel in the carburized condition exhibited longer fatigue life as well as better wear resistance. Nevertheless, [10] compared the contact-fatigue behavior of laser hardened gears with that of carbon case hardened. Results showed that the contact-fatigue strength limit of laser hardened gears is 92% of that of carbon case hardened ones. The reason of this reduction is that the former treatment by laser, possessed too much retained austenite than the later. Surface properties; i.e. wear resistance and contact-fatigue properties of bearing steels have been improved by factors of between 2 to 9, and between 2 to 7, respectively. This was realized by diffusion-chromizing [11]. Shot peening is a process frequently used after case hardening in order to create compressive residual stresses, thus extending fatigue life. This result was achieved [12] with SAE 9245 steel, which has been quenched from 850 C and then tempered. Comparison between shot-peened-heat treated specimens and unpeened ones showed an improvement of 30 % in fatigue life for shot peened specimens. Shot peening was

15

applied again [13] on 4340 steel having a machine-like scratch. Similarly, it was found that the compressive stresses produced due to shot peening has reduced the effect of scratches in which it has increased the fatigue life than that with scratch but un-peened. 4.2. Characteristics of Fatigue Fracture Surfaces [14] Typical fracture surfaces in mechanical components that were subjected to fatigue loads are shown in Fig.1. One characteristic feature of the surface morphology which is evident in both macrographs is the flat, smooth region of the surface exhibiting beach marks (also called clamshell marks). This part represents the portion of the fracture surface over which the crack grew in a stable, slow mode. The rougher regions, showing evidence of large plastic deformation, is the final fracture area through which the crack progressed in an unstable mode.

Fig. 1: Typical fatigue failures in steel components. The beach marks may form concentric rings that point toward the areas of initiation. The origin of the fatigue crack may be more or less distinct. In some cases a defect may be identified as the origin of the crack, in other cases there is no apparent reason why the crack should start at a particular point in a fracture surface. If the critical section is at a high stress concentration fatigue initiation may occur at many points, in contrast to the

16

case of unnotched parts where the crack usually grows from one point only see Fig. 2. While the presence of any defects at the origin may indicate the cause of the fatigue failure, the crack propagation area may yield some information regarding the magnitude of the fatigue loads and also about the variation in the loading pattern. Firstly, the relative magnitude of the areas of slow-growth and final fracture regions give an indication of the maximum stresses and the fracture toughness of the material. Thus, a large final fracture area for a given material indicates a high maximum load, whereas a small area indicates that the load was lower at fracture. Similarly, for a fixed maximum stress, the relative area corresponding to slow crack growth increases with the fracture toughness of the material (or with the tensile strength if the final fracture is a fully ductile overload fracture).

17

Fig. 2: Fracture surfaces at high stress and low stress (schematic) [14] Beach marks are formed when the crack grows intermittently and at different rates during random variations in the loading pattern under the influence of a changing corrosive environment. Beach marks are therefore not observed in the surfaces of fatigue specimens tested under constant amplitude loading conditions without any startstop periods. The average crack growth is of the order of a few millimetres per million cycles in high cycle fatigue, and it is clear that the distance between bands in the beach marks are not a measure of the rate of crack advance per load cycle. However,

18

examination by electron microscope at magnifications between 1,000x and 30,000x may reveal characteristic surface ripples called fatigue striations, see Fig. 3.

Fig. 3: Striations in an aluminum alloy. [14] Although somewhat similar in appearance, these lines are not the beach marks described above as one beach mark may contain thousands of striations. During constant amplitude fatigue loading at relatively high growth rates in ductile material such as stainless steels and aluminum alloys the striation spacing represents the crack advancement per load cycle. However, in low stress, high cycle fatigue where the striation spacing is less than one atomic spacing (- 2.5 x 10-8m) per cycle. Under these conditions the crack does not advance simultaneously along the crack front, growth occurring instead only along some portions during a few cycles, then arrests while growth occurs along other segments. Striations are not seen if the crack grows by other mechanisms such as microvoid coalescence or, in brittle materials, microclevage. In structural steels the crack can propagate by all three mechanism, and striations may be difficult to observe. Fig. 4 shows an example of beach marks and striations in the

19

fracture originating at a large defect in a welded C-Mn steel with a yield strength of about 360MPa.

Fig. 4: Fatigue failures in the Alexander L Kielland platform. [14] 4.3. Nature of the Fatigue Process [14] From the description of the characteristics of fatigue fracture surfaces, three stages in the fatigue process may be identified: Stage I: Crack initiation Stage II: Propagation of one dominant crack Stage III: Final fracture Fatigue cracking in metals is always associated with the accumulation of irreversible plastic strain. The crack process which is discussed in the following applies to smooth specimens made of ductile materials.

20

In high cycle fatigue the maximum stress in cyclic loads that eventually cause fatigue failure may be well below the elastic limit of the material, and large scale plastic deformation does not occur. However, at a free surface plastic strains may accumulate as a result of dislocation movements. Dislocations are line defects in the lattice structure which can move and multiply under the action of shear stresses, leaving a permanent deformation. Dislocation mobility and hence the amount of deformation (or slip) is greater at a free surface than in the interior of crystalline materials due to lack of constraint from grain boundaries. Grains in polycrystalline structural metals are individually oriented in a random manner. Each grain, however, has an ordered atomic structure giving rise to directional properties. Deformation for example, takes place on crystallographic planes of easy slip along which dislocations can move more easily than other planes. Since slip is controlled primarily by shear stress, slip deformation takes place along crystallographic planes that are orientated close to 45 to the tensile stress direction. The results of such deformation are atomic planes sliding relative to each other, resulting in a roughening of the surface in slip bands. During further cycling slip band deformation is intensified at the surface and extending into the interior of the grain, resulting in so-called persistent slip bands, (PSB's). The name originated from the observation in early studies of fatigue that slip band would reappear - "persist" - at the same location after a thin surface layer was removed by elastopolishing. The accumulation of local plastic flow result in surface ridges and troughs called extrusions and intrusions, respectively, Fig. 5. The cohesion between the layers in slip band is weakened by oxidation of fresh surfaces and hardening of the strained material. At some point in this process small cracks develop in the intrusions.

21

Fig. 5: Slip band with extrusions and intrusions formed on the surface of a grain subjected to cyclic stress. Crack nucleation at intrusion. [14] These microcracks grow along slip planes, ie. a shear stress driven process. Growth in the shear mode, called stage I crack growth extends over a few grains. During continued cycling the microcracks in different grains coalesce resulting in one or a few dominating cracks. The stress field associated with the dominating crack cause further growth under the primary action of maximum principal stress; this is called stage II growth. The crack path is now essentially perpendicular to the tensile stress axis. Crack advancement is, however, still influenced by the crystallographic orientation of the grains and the crack grows in a zigzag path along slip planes and cleavage planes from grain to grain, see Fig. 6.

22

Fig. 6: Stage I and stage II crack growth at polycrystalline material. [14] Most fatigue cracks advance across grain boundaries as indicated in Fig. 6, ie. in a transcrystalline mode. However, at high temperatures or in a corrosive environment, grain boundaries may become weaker than the grain matrix, resulting in intercrystalline crack growth. The fracture surface created by stage II crack growth are in ductile metals characterized by striations whose density and width can be related to the applied stress level. Since crack nucleation is related to the magnitude of stress, any stress concentration in the form of external or internal surface flaws can marked reduce fatigue life, in particular when the initiation phase occupies a significant portion of the total life. Thus a part with a smooth, polished surface generally has a higher fatigue strength than one with a rough surface. Crack initiation can also be facilitated by inclusions, which act as internal stress raisers. In ductile materials slip band deformations at inclusions are higher than elsewhere and fatigue cracks may initiate here unless other stress raisers dominate.

23

In high strength materials, notably steels and aluminium alloys, a different initiation mechanism is often observed. In such materials, which are highly resistant to slip deformation, the interface between the matrix and inclusion may be relatively weak, and cracks will start here if decohesion occurs at the inclusion surface, aided by the increased stress/strain field around the inclusion. Fig. 7 shows small fatigue cracks originating at inclusions in a high strength steel. Alternatively, a hard brittle inclusion may break and a fatigue crack may initiate at the edges of the cleavage fracture.

Fig. 7: Fatigue crack initiation at an inclusion in a high strength steel alloy. [14] From the discussion above it is evidently not possible to make a clear distinction between crack nucleation and stage I growth. "Crack initiation" is thus a rather imprecise term used to describe a series of events leading to stage II crack. Although the initiation stage includes some crack growth, the small scale of the crack compared with microstructural dimensions such as grain size invalidates a fracture mechanics based analysis of this growth phase. Instead, local stresses and strains are commonly related to material constants in prediction models used to estimate the length of stage I. The material constants are normally obtained from tests on smooth specimens subjected to stress or strain controlled cycling.

24

The initiation process usually results in a number of micro-cracks that may grow more or less independently until one crack becomes dominant through a coalescence process at the microcracks start to interact. Under steady fatigue loading this crack grows slowly, but starts to accelerate when the reduction of the cross-section increases the local stress field near the crack front. Final failure occurs as an unstable fracture when the remaining area is too small to support the load. These stages in the fatigue process can in many cases be related to distinctive features of the fracture surface of components that have failed under fluctuating loads, the presence of these features can therefore be used to identify fatigue as the probable cause of failure.

25

5. EXPERIMENTAL PROCEDURES
In the present work, three grades of steel namely; AISI/ SAE 1040 [Carbon Steel], 4140 [Chromium Molybdenum steel] and 4340 [Nickel Chromium Molybdenum steel have been selected to be investigated in the high-cycle fatigue regime under conditions of rotating bending. These steels were cold-drawn round bars with diameters 25, 24 and 22 mm. Types of steel selected are widely used in industry and are mainly characterized by durability which is perhaps the most significant attribute an industry can possess. Steel 1040 is normally used in axels, crank shafts and gears, whereas, steel 4140 is used connecting rods, hydraulic rams and spindles. Steels 4340 is used in air craft landing gears, power transmission gears and shafts. The chemical composition, in wt.% and the mechanical properties of these steels are shown in tables 1 and 2 respectively. Table 1: Chemical composition of materials (wt %). [as supplied by the manufacturer] Steel Grade AISI/SAE 1040 4140 4340 C 0.41 0.42 0.36 Mn 0.6 0.75 0.25 P 0.04 0.04 0.04 S 0.05 0.04 0.04 Si 0.3 0.25 0.25 Cr 1.05 1.4 Mo 0.2 0.2 Ni 1.4

Table 2: Mechanical properties of materials tested. [tested according to ASTM E8-00] Steel Grade AISI/SAE 1040 4140 4340 UTS MPa 580 655 745 Yield strength MPa 489 417 473 Elongation % 11 26 22 Hardness Vickers 300 g 220 270 255 Impact toughness J 22 55 51

26

The steel samples were heat treated according to the steps summarized in Table 3. Table 3: Heat Treatment procedures applied to the three steel types.
Steel Type Quenching and Tempering 1. Austenitize at 850 C and hold for 1 hour. 2. Quench rapidly in water with vigorous stirring. 3. Temper by reheating at 450 C and hold for 2 hours. 4. Cool in air. 1. Austenitize at 850 C and hold for 1 hour. 2. Quench rapidly in oil with stirring. 3. Temper by reheating at 550 C and hold for 2 hours. 4. Cool in still air. 1. Austenitize at 850 C and hold for 1 hour. 2. Quench rapidly in oil with stirring. 3. Temper by reheating at 550 C and hold for 2 hours. 4. Cool in still air. Heat Treatment Normalizing 1. Austenitize at 850 C and hold for 1 hour. 2. Cool in still air. Annealing 1. Austenitize at 850 C and hold for 1 hour for reaching uniform temperature throughout the section. 2. Cool in furnace.

Steel 1040

Steel 4140

1. Austenitize at 880 C and hold for 1 hour. 2. Cool in still air.

1. Austenitize at 830 C and hold for 1 hour for reaching uniform temperature throughout the section. 2. Cool in furnace. 1. Austenitize at 830 C and hold for 1 hour for reaching uniform temperature throughout the section. 2. Cool in furnace.

Steel 4340

1. Austenitize at 880 C and hold for 1 hour. 2. Cool in still air.

The smooth round-stepped bar specimens used to evaluate the fatigue life had a gage length of 109.5 mm with a cross-sectional diameter of 8 mm, as shown in Fig. 8, with a 1 mm radius of curvature at the critical section.

Fig. 8: Fatigue specimen geometry, RA= 1mm. Dimensions in mm

27

These specimens were machined close to their size before heat treatment, and then polished with abrasive papers number 1000 after the heat treatment. The fatigue tests were performed at room temperature using a rotatingbending fatigue testing machine at a rotational speed of 3000 rpm. The test was with a constant amplitude sinusoidal loading pattern at a frequency of 50 Hz and a stress ratio of R= - 1 (tension compression). The occurrence of fracture was controlled by a limiting switch and the number of cycles at fracture recorded by a built-in counter. The maximum stress amplitude was calculated according to the relation: max= Mc/ I where, M is the bending moment at the specimens critical section, c is the maximum distance from specimens center line and I is the moment of inertia of a round cross section: r4 / 4.

28

6. RESULTS AND DISCSSION


6.1. Effect of Heat treatment on Hardness Fig. 9 shows the effect of different heat treatment on the hardness of the steels studied in this work.

600
1040

Hardness Vickers

500 400 300 200 100 0

4340

4140

4340 1040

4140

4340 1040 4140

1 QT

2 N Heat Treatment Type

3 A

Fig. 9: Hardness values for three types of steel with different heat treatments QT: Quenched and Tempered N: Normalized A: Annealed 6.2. Effect of Heat Treatment on Mechanical Properties Table 4 presents the effect of different heat treatment on the mechanical properties of the steels studied in this work. Table 4: Mechanical properties of three types of steel measured after different heat treatments.
Steel Grade AISI/ SAE 1040 4140 4340 Ultimate Tensile Strength MPa QT N A
1196 1070 1133 732 1196 1968 559 661 708

Yield Strength MPa QT


929 944 1039

Elongation % QT
21 30 29

Hardness Vickers 300 g A


50 37 41

Impact Toughness J QT
147 130 129

N
523 992 1669

A
330 362 441

N
18 21 27

QT
455 378 367

N
256 421 390

A
180 219 210

N
137 117 121

A
116 108 110

QT: Quenched and Tempered

N: Normalized

A: Annealed

29

6.3. Effect of Various Heat Treatments on Fatigue Life of 4340 Steel


700 600 500 400 300 200 1.00E+02

stress amplitude, MPa

N A

Q&T

1.00E+03

1.00E+04

1.00E+05

1.00E+06

number of cycles to failure, Nf

Fig. 10: S-N curve for 4340 steel with different heat treatments Q&T: Quenched and Tempered N: Normalized A: Annealed

30

Quenching and tempering treatment gives the longest fatigue life when compared to normalizing and annealing treatments. Microstructural study showed that the quenched and tempered microstructure is predominantly tempered martensite, Fig. 11: a, whereas, normalizing gave bainite with occasional retained austenite. Annealing led to a soft structure; Fig. 11: c, which contains ferrite [white] and pearlite [black]. The homogeneous tempered martensite has the highest fatigue properties with respect to the other microstructures, in spite of having moderate tensile and yield strengths as in Table 4. This is can be understood based on the fact that homogeneous structures results in homogeneous slip deformations so that local concentration of plastic deformations are avoided, thus causing dislocation cross-slip more difficult [15]. It is also understandable that a microstructure consisting phases with contrasting difference in deformability would promote slip localization which contributes detrimentally to each stage of fatigue [16]. This analysis is seen to be against the fact that fatigue properties at high hardness are extremely sensitive to residual stresses and inclusions. Another reason for this superior fatigue property may be due to the oil quenching that occurred with this treatment. With such alloy steels, oil quenching results into martensite formation with its accompanying volume increase [17]. This may lead to residual compressive stresses which relatively makes the crack spend more energy to grow and consequently hinders an initiated crack to propagate.

31

a. Quenched and Tempered

b. Normalized

c. Annealed Fig. 11: Steel 4340, Nital etched, x 1200 a. Quenched and Tempered b. Normalized c. Annealed

32

6.4. Effect of Various Heat Treatments on Fatigue Life of 1040 Steel


700 600 500 400 300 200 1.00E+02

stress amplitude, MPa

N A

Q&T

1.00E+03

1.00E+04

1.00E+05

1.00E+06

number of cycles to failure, Nf

Fig. 12: S-N curve for 1040 steel with different heat treatments Q&T: Quenched and Tempered N: Normalized A: Annealed

33

Similarly, and stated above, quenching and tempering treatment resulted in longest fatigue life when compared to normalizing and annealing treatments respectively, Fig. 12. This is seen to be due to two reasons; first is the size and density of inclusions present [19] and second is the microstructure type in within which crack propagates during the entire fatigue life. Fig. 13: a, b and c shows at low magnification, the size, distribution and density of inclusions found in steel 1040. The annealed condition revealed the biggest amount and largest size of inclusions, whereas, and to less extent, normalized and quenched and tempered respectively.

a. Quenched and Tempered

b. Normalized

34

c. Annealed Fig. 13: Steel 1040, unetched, x 120 a. Quenched and Tempered b. Normalized c. Annealed

The microstructural study revealed that the quenched and tempered structure is homogeneous fine-laths of tempered martensite, Fig. 14: a, whereas, the normalized structure consists of grains of ferrite [white] and pearlite [black], Fig. 14: b. Annealed structure was coarse-grains of ferrite and pearlite Fig. 14: c.

a. Quenched and Tempered

35

b. Normalized

c. Annealed Fig. 14: Microstructure of steel 1040, Nital etched, x1200 a. Quenched and Tempered b. Normalized c. Annealed The homogeneous tempered martensite has the highest fatigue properties with respect to the other microstructures, in spite of having moderate tensile and yield strengths as in Table 4. This is can be understood based on the fact that homogeneous structures results in homogeneous slip deformations so that local concentration of plastic deformations are avoided, thus causing dislocation cross-slip more difficult [15]. This analysis is seen to be against the fact that fatigue properties at high hardness are extremely sensitive to residual stresses and inclusions. Another reason for this superior fatigue property may

36

be due to the vigorous quenching [water with agitation] that occurred with this treatment. Quenching results into martensite formation with its accompanying volume increase [17]. This leads to residual compressive stresses which relatively makes the crack spend more energy to grow and consequently hinders an initiated crack to propagate. Annealing treatment results into lowest fatigue properties due to the soft ferriticpearlitic structure in which ferrite grains dominate, as seen in Fig. 14: c. Within the stages of crack initiation and propagation, the soft ferrite grains are deformed heavily by forming fine and uniform slip bands [18] thus encouraging fatigue cracks to initiate at slip bands rather than at the ferrite-pearlite interface, although the notch effect of carbide lamellae in pearlite may be considered as another reason for fatigue strength reduction.

37

6.5. Effect of Various Heat Treatments on Fatigue Life of 4140 Steel


700 600 500 400 300 200 1.00E+02

stress amplitude, MPa

Q&T

1.00E+03

1.00E+04

1.00E+05

1.00E+06

number of cycles to failure, Nf

Fig. 15: S-N curve for 4140 steel with different heat treatments Q&T: Quenched and Tempered N: Normalized A: Annealed

38

Similarly, and as seen before, the quenched and tempered treatment gave the longest fatigue life when compared to N and A treatments, Fig. 15. This can also be attributed to the amount of inclusions existing in each type of treatment. Fig. 16 shows that quenched and tempered specimen possesses the minimum amount of inclusions with respect to size, distribution and density.

a. Quenched and Tempered

b. Normalized

39

c. Annealed Fig. 16: Steel 4140 with various heat treatments, unetched. x 120 a. Quenched and Tempered b. Normalized c. Annealed 6.6. S-N Curve Equations The power law Basquin relationship [20], = Constant Nf n, represents the equation of S- N curve, where: is the stress amplitude in MPa, Nf is the number of cycles to failure and n is the equations exponent. Table 5 presents the values of the constant and the exponent n for the S- N curves obtained in this work for all steel types and their heat treatments. The constant is the stress value at one cycle and the exponent n is the slope of the curve which means that longer fatigue lives occur with lower stress amplitudes.

Table 5: The Constant and exponent n values in Basquin relationship.


Steel Type [Heat Treatment Type] 1040 [Annealed] 1040 [Normalized] 1040 [Quenched and Tempered] 4140 [Annealed] 4140 [Normalized] 4140 [Quenched and Tempered] 4340 [Annealed] 4340 [Normalized] 4340 [Quenched and Tempered] Constant 2424 4056 5599 2575 6669 8626 2606 43074 4098 N 0.2256 0.2404 0.2477 0.2107 0.2656 0.2764 0.1841 0.4214 0.1957

40

It can be noted from Table 5 that annealing, normalizing and quenching and tempering treatments have similar effect on three types of steel investigated with respect to the values of the equations constant and exponent with the exception of steel 4340 in the normalized condition. Moving from annealing to normalizing then to quenching and tempering is accompanied by increasing both the constant and exponent values. This is in agreement with the results discussed above where quenching and tempering treatment gives the longest fatigue life when compared to normalizing and annealing respectively. The relatively high values of the constant and exponent n of steel 4340 in the normalized condition may be attributed the high hardness [390 HV] that resulted from normalizing treatment when compared to quenching and tempering [367 HV] and annealing [210 HV]. The crack initiation and propagation process is normally faster in hard microstructures than that in relatively softer ones thus causing shorter fatigue life.

41

6.7. Effect of Quenching and Tempering on Fatigue Life of Different Steels


700

stress amplitude, MPa

4140
600 500 400 300 200 1.00E+02

4340 1040

1.00E+03

1.00E+04 cycles to failure, Nf

1.00E+05

1.00E+06

Fig. 17: S-N curves for three types of steels in the quenched and tempered condition

42

Fig. 17 shows the effect of quenching and tempering treatment on the steels studied. Steel 4340 gave the longest fatigue life when compared to 4140 and 1040 respectively. Size, distribution and density of slag inclusions are thought to be one reason for this difference. The detrimental effect of slag inclusions on fatigue life has been mentioned earlier. Fig. 18 shows slag inclusions are biggest in amount with steel 1040 and to a less extent with 4140 and 4340 steels.

b. Steel 4340

c. Steel 4140

43

d. Steel 1040 Fig. 18: Slag Inclusions in the Quenched and Tempered for different steel types, unetched, x120 a. steel 4340 b. steel 4140 c. steel 1040 Another reason for the relatively longer fatigue life of steel 4340 than that of steels 4140 and 1040 respectively is the microstructure type. Fig. 19 shows that the quenched and tempered microstructures are entirely lath-martensite. However, steel 4340 and 4140 had similar martensite [coarse] with occasional retained austenite [white areas], whereas, steel 1040 has very fine martensite. This is due to the difference in the chemical analysis of steel types studied and also due to the cooling medium used in quenching these steels. Severe quenching using water was applied with 1040, whereas, oil with 4340 and 4140 steels. Steel 4340 possesses the highest yield strength [1039 MPa] than steels 4140 [944 MPa] and 1040 [929 MPa]. Hence, steel 4340 have better fatigue properties based on the fact that higher yield point materials have higher fatigue strengths because they possess high resistance to crack initiation. In addition, slip homogeneity is more encouraged in a primarily single phase microstructure (i.e., tempered martensite) than otherwise (i.e., retained austenite and tempered martensite) [16]. Moreover, the needle-shaped cementite distributed within the tempered martensite

44

with a roughly definite orientation, as well as along the lath boundary, could act as obstacles to dislocation slip and thus tends to reduce fatigue properties. [16] This is in agreement with hardness testing results, which showed that steel 1040 is the hardest material [455 HV] compared to steels 4140 [378 HV] and 4340 [362 HV]. The harder the material is, the more structure sensitive it is to crack initiation and propagation.

a. Steel 4340

b. Steel 4140

45

c. Steel 1040 Fig. 19: Optical micrographs of quenched and tempered steels.x1200 a. steel 4340 b. steel 4140 c. steel 1040 6.8. Characteristics of Fatigue-Fractured Surfaces Subjected to Different Heat Treatments 6.8.1. Annealing With almost all annealed specimens several remarks can be deduced from photo macrographs taken at ten times magnification. The fractured surface is characterized by several dominating cracks that has initiated at the specimens surface, at which stresses are maximum, and propagated radially within the cross-section [light zone], sometimes called slow-crack growth zone, Fig. 20 This propagation continues until the remaining area of the specimens cross-section is too small to support the load [dark zone], Fig. 20. The surface of this dark zone, sometimes called instantaneous zone, which suddenly failed, is characterized as relatively rough or fibrous and containing voids, inclusions and discontinuities. The fibrous surface indicates that large plastic deformations have occurred. This explains why the annealed condition gave the lowest fatigue life irrespective to the steel type.

46

a. Steel 1040

b. Steel 4140

c. Steel 4340

Fig. 20: Macrographs of fractured surfaces of annealed specimens. x 10 6.8.2. Normalizing On the other hand, the effect of normalizing treatment on the fractured surface can be seen in Fig. 21. The outer surface of the specimens is characterized by relatively much less dominating cracks than in the annealed condition, however, still there are some radial cracks exist.

a. Steel 1040

b. Steel 4140

c. Steel 4340

Fig. 21: Macrographs of fractured surfaces of normalized specimens. x 10 It can be also noted from Fig. 21 that the slow-crack growth zone became slightly bigger on the expense of the final fractured zone, thus indicating that normalizing treatment results in longer fatigue lives than annealing. This result can be manifested

47

from the smoothness of the slow-crack growth zone, where normalizing resulted in a smoother surface than annealing. This reflects that crack growth stage was slower; i.e. longer, and more stable than annealing.

a. Steel 4140 normalized

b. Steel 4140 annealed

Fig. 21: Macrographs comparing slow-crack growth zone between normalizing and annealing for the same steel type. x 10

6.8.3. Quenching and tempering

Fig. 22 shows the fractured surface of the quenched and tempered treatment for different steel types. It can be noted that the size of the cracks normally found at the outer surface of the specimen became much smaller than that with the normalizing and annealing. In addition, the slow-crack growth zone is now smoother and larger, thus indicating longer fatigue lives when compared to the other two treatments.

48

a. Steel 1040

b. Steel 4140 Fig. 22: Macrographs of fractured surfaces of quenched and tempered specimens. x 10

c. Steel 4340

6.9. Impact Toughness Fig. 23 shows the impact toughness of the three steel types when subjected to different heat treatments. Steel 1040, with all treatments, is the toughest type of steel compared to steels 4140 and 4340. Whereas, an ignorable difference in toughness between 4140 and 4340 steels. In addition, and similar to fatigue results, quenching and tempering treatment resulted in highest toughness results, whatever the steel type is. These results are in good agreement with fractured surface studied above, which indicated that the larger the slow-crack growth zone is [occurring with quenching and tempering treatment], the longer the fatigue life [14].

49

160 140 120 100 80 60 40 20 0

Impact Toughness, J

1040 4340

4140 4340

1040 4140 4340 1040 4140

Q &1T

2 N Heat Treatment

3A

Fig. 23: Impact toughness test results for three types of steels subjected to different heat treatments Q&T= Quenching and Tempering N: Normalizing A: Annealing

50

7. CONCLUSIONS
The rotating bending fatigue tests have been carried out to evaluate the effect of different heat treatments on three types of steels; namely AISI / SAE 1040, 4140 and 4340. The following conclusions are drawn: 1. Quenching and tempering treatments gave the longest fatigue life compared to normalizing and annealing respectively. This equally applies for the three types of steels investigated. 2. Steel 4340 gave the highest fatigue life with respect to the other steels. 3. Steel 1040 was the toughest type of steel with respect to other steels whatever the type of heat treatment is. 4. Microstructural and Fractographic investigations are very important tools that should be used in failure analysis. 5. Slag inclusions play an important role in the fatigue life of a material.

51

8. REFERENCES
1. Chitty, J.A.; Pertuz, A.; Hintermann, H.; Puchi, E.S, Influence of electroless nickel-phosphorus deposits on the corrosion-fatigue life of notched and un-notched samples of an AISI 1045 steel, Journal of Materials Engineering and Performance, Volume: 8, Issue: n 1, Pages: p 83-86, 1999. 2. Garces, Y.; Sanchez, H.; Berrios, J.; Pertuz, A.; Chitty, J.; Hintermann, H.; Puchi, E.S., Fatigue behavior of a quenched and tempered AISI 4340 steel coated with an electroless Ni-P deposit, Proceedings of the 1999 26th International Conference on Metallurgic Coatings and Thin Films, San Diego, CA, USA, Apr 12- 15 1999. 3. Braza, J.F, Rolling contact fatigue and sliding wear performance of ferritic nitrocarburized M-50 steel, Tribology Transactions, v 35, Issue: n 1, Pages: p 8997, Jan. 1992. 4. Yang, Z.G.; Yao, G.; Li, G.Y.; Li, S.X.; Chu, Z.M.; Hui, W.J.; Dong, H.; Weng, Y.Q., The effect of inclusions on the fatigue behavior of fine-grained high strength 42CrMoVNb steel, International Journal of Fatigue, v 26, Issue: n 9, Pages: p 959966, September 2004. 5. Mediratta, S.R.; Ramaswamy, V.; Singh, V.; Rao, P. Rama, Low cycle fatigue of dual phase steels produced by different cooling rates and a ferrite-pearlite steel, Scripta Metallurgica et Materialia, v 24, Issue: n 4, Pages: p 793-797, Apr. 1990 6. Fly, David E.; Black, J.T.; Singleton, Ben, Low power laser heat treatment to improve fatigue life of low carbon steel, Journal of Laser Applications, v 8, Issue: n 2, Pages: p 89-93, Apr. 1996. 7. Chen, Jianqiao; Takezono, Shigeo; Li, Guangxia; Tanaka, Toshiyuki, Effect of laser cladding on fatigue strength of an alloy steel, Journal of the Society of Materials Science, Japan, v 44, Issue: n 498, Pages: p 343-347, Mar. 1995. 8. Da Silva, V.F.; Canale, L.F.; Spinelli, D.; Bose-Filho, W.W.; Crnkovic, O.R., Influence of retained austenite on short fatigue crack growth and wear resistance of case carburized steel, Journal of Materials Engineering and Performance, v 8, Issue: n 5, Pages: p 543-548, Oct. 1999. 9. Furumura, Kyozaburo; Murakami, Yasuo; Abe, Tsutomu, Case-hardening medium carbon steel for tough and long life bearing under severe lubrication conditions, ASTM Special Technical Publication, v 1327, Pages: p 293-306, Apr. 1998. 10. Zhang, H.; Shi, Y.; Xu, C.Y.; Kutsuna, M, Comparison of contact fatigue strength of carbon case hardening and laser hardening of gears, Surface Engineering, v 20, Issue: n 2, Pages: p 117-120, April 2004. 11. Dubinin, G.N.; Oganesyan, G.L., Effect of a diffusion layer on the properties of bearing steel, Metal Science and Heat Treatment (English Translation of Metallovedenie i Termicheskaya Obrabotka Metallov), v 32, Issue: n 1-2, Jul. 1990. 12. Tekeli, S., Enhancement of fatigue strength of SAE 9245 steel by shot peening, Materials Letters, v 57, Issue: n 3, Pages: p 604-608, December 2002. 13. Everett Jr., R.A.; Newman Jr., J.C.; Phillips, E.P, The effects of a machining-like scratch on the fatigue life of 4340 steel, Journal of the American Helicopter Society, v 47, Issue: n 3, Pages: p 151-155, July 2002. 14. Electronic Web Site: http://www.kuleuven.ac.be 15. P. Talukdar, S.K. Sen, A.K. Ghosh, Effect of Fatigue damage on the dynamic fracture toughnes of En-8-Grade Steel, Metallurgical and Materials Transactions A, volume 32A, October 2001- p. 2547

52

16. C.S. Lee, K.A. Lee, D.M. Li, S.J. Yoo, W.J. Nam, Microstructural influence on fatigue properties of high-strength spring steel, Materials Science and engineering, A241 (1998) pp. 30-37. 17. V.F. Siva, L.F. Canale, D. Spinelli, W.W. Bose-Filho, and O.R. Crnkovic, Influence of retained austenite on short fatigue crack growth and wear resistance of case carburized steel, Journal of Materials Engineering and Performance, Volume 8 (5) 1999- p. 543 18. Z.G. Wang and S.H.Al, Fatigue of Martensitr-Ferrite High Strength Low-alloy dual Phase steels, ISIJ International, Volume 39 (1999), No. 8, pp. 747- 759. 19. Z.G. Yang, G. Yao, G.Y. Li, S.X. Li, Z.M. Chu, W.J. Hui, H. Dong, Y.Q. Weng, The effect of inclusions on the fatigue behaviour of fine-grained high strength 42CrMoVNb steel, International journal of fatigue 26 (2004) 959-966. 20. S. Sankaran , V. Subramanya Sarma , K.A. Padmanabhan , G. Jaeger b, A. Koethe, High cycle fatigue behaviour of a multiphase microalloyed medium carbon steel: a comparison between ferritepearlite and tempered martensite microstructures, Materials Science and Engineering A362 (2003) 249256.

53

También podría gustarte