You are on page 1of 7

International Journal of Fatigue 26 (2004) 683689 www.elsevier.

com/locate/ijfatigue

A fatigue life model for 5% chrome work roll steel under multiaxial loading
K.S. Kim , K.M. Nam, G.J. Kwak, S.M. Hwang
Department of Mechanical Engineering, Pohang University of Science and Technology, San 31 Hyoja-dong, Nam-gu, Pohang 790-784, South Korea Received 5 August 2003; received in revised form 18 September 2003; accepted 20 November 2003

Abstract The fatigue behavior of 5% chrome steel heat-treated for wear resistance has been investigated under axialtorsional loading. This material exhibits brittle fracture under monotonic and cyclic loading. The preferred site for crack initiation appears to be carbide clusters on or near the surface. Crack propagation initially progressed in a transgranular mode followed by a mixed transgranularintergranular mode at a later stage. A parameter given in terms of the maximum normal stress range and the hydrostatic stress range is found to correlate fatigue lives reasonably well. This parameter correctly predicts the experimental trend that in-phase loading is more damaging than out-of-phase loading under a given ratio of axial/shear stress amplitudes. Models for tensile and compressive mean stress eects have also been proposed based on the uniaxial test results. # 2003 Elsevier Ltd. All rights reserved.
Keywords: Multiaxial fatigue; Normal fracture; Mean stress eect; 5% Chrome steel; Work roll

1. Introduction Heat-treated 5% chrome steel is widely used for work rolls in cold rolling operations. Work rolls are subjected to surface damage of various types. Among these are normal wear, bruising and spalling. Micro-cracks are commonly observed in ultrasonic and eddy current testing during periodic dressing of the roll surface. Cracks may be initiated by thermal shocks [13] from frictional skidding between the roll and work piece, or from excessive local deformation due to pinching and cobbling of the incoming strip. The local heat buildup may result in material softening and crack initiation upon sudden cooling. Once initiated, cracks can propagate under repeated contact loads, and accidental spalling of the roll surface may entail. Another type of surface damage of a work roll is fatigue damage due to repeated contact loads. The material in the contact zone is subjected to complex, multiaxial, compressive stresses, which can initiate fatigue
Corresponding author. Tel.: +82-54-279-2182; fax: +82-54-2795899. E-mail address: illini@postech.ac.kr (K.S. Kim).

cracks and cause surface damage. Contact fatigue models have been reviewed by Tallian [4], Alfredsson and Olsson [5] recently. The immediate area outside the contact zone in the roll undergoes tension, which may become large enough for fatigue considerations in adverse circumstances, such as the presence of sharp corners at the contact edge or material defects. Salehizadeh and Sakas analysis [6] shows that the matrix material around hard inclusions in the contact zone experiences tensile stresses if the contact load is high enough to produce plastic strain in the matrix. The material under investigation can be deformed plastically under highly concentrated contact loads. The material is found to be much weaker in tension than in compression. Therefore, studies on tensile fatigue fracture, as well as compressive fatigue resistance under contact load, would be of considerable practical interest. The purpose of this paper is to investigate the fatigue behavior of heat-treated, 5% chrome steel under uniaxial, torsional and combined axialtorsional loading conditions that lead to tensile fatigue fracture. A fatigue model will be proposed suitable for the material under the loading conditions employed. Some uniaxial tests

0142-1123/$ - see front matter # 2003 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijfatigue.2003.11.005

684

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

Table 1 Chemical composition of 5% chrome steel Element wt% C 0.88 Si 0.33 Mn 0.35 P 0.009 S 0.001 Ni 0.11 Cr 4.97 Mo 0.41 Cu 0.33

will also be conducted in the presence of mean stress, and a model will be proposed for the mean stress eect. The results obtained in this work will be useful for studying the multiaxial fatigue behavior of other similar hard steels such as tool steels and high-speed steels.

2. Material and specimen The chemical composition of 5% chrome steel under consideration is given in Table 1. Specimens used in this study are solid specimens with a round crosssection. The gage section had a diameter of 6 mm and a length of 20 mm. The specimens were machined from forged material that has undergone electrical remelting and spheroidizing. Prior to heat treatment initial machining was done such that the diameters at the gage section and at the grip area were 0.5 mm larger than the nal dimension. The specimen was then austev v nized at 980 C for 20 min and oil quenched at 60 C. v Then, the specimen was tempered for 3 h at 130 C and cooled in air. The gage section of the heat-treated specimen was ground at a very slow rate to the nal size to ensure that heat generated in the grinding process would not aect hardness. Finally, the gage section was polished with alumina powder. After this step, the specimen had hardness in the range of Rc 6465. The microstructure of the material is primarily tempered martensite matrix with carbides (Fe,Cr)7C3 [7,8] dispersed in the matrix and on the prior austenite grain boundary, as shown in Fig. 1. Most of the grains were

observed in the size range of 1015 lm. The average carbide diameter was 0.52 lm, with some having diameters as large as 2 lm. The monotonic stressstrain curve of this material is asymmetric in tension and compression. The compression test was conducted in a nonstandard way using a fatigue specimen, and terminated before failure occurred for safety reasons. It was found that this material is much stronger in compression than in tension (see Fig. 2). The mechanical properties obtained from these tests are: Youngs modulus 197 GPa; 0.2% tensile yield strength 900 MPa; tensile strength 1400 MPa; elongation 1.3%; 0.2% compressive yield strength 2000 MPa; compressive strength >2500 MPa.

3. Experiment Fatigue tests were carried out with an axialtorsional servo-hydraulic Instron machine at room temperature. The waveform utilized was triangular, and frequency varied from 0.5 to 5 Hz, with higher frequencies applied to lower amplitude tests. Failure was dened as a complete separation of the specimen. All tests were carried out under load control. It was unnecessary to measure strain because the loaddisplacement response

Fig. 1. Microstructure of test material.

Fig. 2.

Monotonic stressstrain curve in tension and compression.

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

685

was approximately linear for all test conditions. Loading conditions for uniaxial tests and torsional tests are given in Table 2. The combined axialtorsional tests are summarized in Table 3. All tests in Table 3 are fully reversed tests. Three phase angles between the axial and shear stresses were investigated under a given ratio of stress amplitudes. It is noted that the phase angles were found somewhat shifted from the intended v v v values of 0 , 45 and 90 (see Table 3), where suces 1, 2, and 3 in the specimen numbers are in the order of increasing phase angles. The phase angle, nevertheless, stayed steady during each test.

Table 3 Loading conditions for combined axialtorsional fatigue tests Specimen no. A1 A2 A3 B1 B2 B3 C1 C2 C3 D1 D2 D3 E1 E3-1 E3-2 F3 G1 Axial stress (MPa) 747 748 735 589 589 594 423 434 430 674 690 680 551 550 549 642 396 Shear stress (MPa) 432 433 432 598 605 597 747 751 745 390 394 390 548 549 543 791 499 Phase angle ( ) 14 58 94 14 58 101 7 50 101 14 72 108 14 101 94 79 14
v

4. Fatigue life prediction The fatigue parameter to be used in this study is given in terms of the maximum normal stress range and the hydrostatic stress range. This parameter has been constructed based on the fact that fracture of a brittle material is dictated by the maximum tensile stress, and on the fact that the parameter must be able to describe the relative experimental trends for dierent loading conditions. The fatigue criterion is given by: Pn Drn k1 Drh r0f 2Nf b ; 1

where Pn is the normal parameter, Drn is the maximum normal stress range, Drh is the hydrostatic stress range,
Table 2 Uniaxial and torsional fatigue test conditions and observed lives Loading type Uniaxial Stress amplitude Mean stress (MPa) (MPa) 950 850 800 700 650 600 600 500 400 350 900 850 800 750 720 900 850 820 800 1000 900 850 800 700 650 0 0 0 0 0 0 300 300 300 300 100 100 100 100 100 300 300 300 300 0 0 0 0 0 0 Life (cycles) 290 1658 4516 24,434 102,032 128,175 3479 65,657 519,592 2,388,905 3422 2320 112,038 308,211 81,438 8486 192,110 45,819 2,615,767 8595 18,454 13,096 95,313 367,012 1,017,325

Uniaxial

Uniaxial

Uniaxial

Torsional

Nf is the number of cycles to failure, k1, b and r0f are material constants to be determined from uniaxial and torsional fatigue test data. The plasticity eect has not been included in this fatigue criterion since the cyclic stressstrain response of the material was elastic for all test conditions. The hydrostatic stress term in the parameter was introduced to model the dierences in fatigue lives between uniaxial and torsional loading with identical maximum normal stress ranges. The hydrostatic stress was entered in various fatigue criteria in dierent ways. Haigh [9] and Sines [10] considered the mean hydrostatic stress as a measure of the mean stress eect in multiaxial fatigue. Fatigue properties were set to vary with hydrostatic stress in the fatigue criteria by Libertiny [11], Davis and Connelly [12], Manson and Halford [13], and Kalluri and Bonacuse [14]. Kakuno and Kawada [15] proposed a criterion given in terms of equivalent stress range, hydrostatic stress range and mean hydrostatic stress. Dang-Van [16] used shear stress and hydrostatic stress ranges in the parameter for high cycle fatigue life prediction. It does not appear, however, that the hydrostatic stress range has been used with normal stress criteria for brittle materials. The computation of fatigue life was carried out using a computer program written for this study. It was assumed that the crack initiates and propagates on the critical plane where the maximum normal stress range occurs. The critical plane was determined using an v increment of 1 in the angle of orientation of the plane v v that varied from 0 to 180 from the specimen axis. The fatigue parameter and life were then evaluated on the critical plane. It is worth noting that a shear fatigue criterion given by Ds sDrn s0f 2Nf b , where Ds is the maximum

686

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

shear stress range, Drn the normal stress range on the Ds plane, s0f , s, b are fatigue properties, was also investigated for the possibility of shear fracture in the very early phase, which is dicult to verify because of the extremely small crack sizes involved. A similar attempt can also be found in an early study by Sines [10] on cast iron. For type 304 stainless steels, which show normal fracture under most loading conditions, a shear fracture parameter of this type, but given in strain, provides a reasonable correlation of life data [1719]. It was found, however, that this parameter does not provide the experimental trend for the current material that in-phase loading is more damaging than out-ofphase loading. The eort was then abandoned.

Fig. 3. sional.

Fractured specimens in fatigue tests: (a) uniaxial, (b) tor-

5. Results and discussion The orientation of fracture planes in uniaxial and torsional tests resembled that of a typical brittle v v material; 90 for uniaxial loading and 45 for torsional loading as shown in Fig. 3. A low magnication SEM

micrograph of the fracture surface (Fig. 4(a)) was taken on a specimen subjected to uniaxial loading of Dr 350 MPa, Rr 0:125. The crack appears to have initiated in the area marked by A and propagated approximately 0.4 mm deep in thumbnail shape before the onset of unstable fracture occurred. SEM micrographs of higher magnication revealed that crack initiation might have taken place at a cluster of subsurface car-

Fig. 4.

SEM micrographs of fracture surface: (a) overall view, (b) site A, (c) site B, (d) site C.

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

687

bide particles (Fig. 4(b)). A similar observation can be found in Mueling et al. [7] for similar steels. The crack propagated in a transgranular mode initially (Fig. 4(c)) which was followed by mixed transgranular and intergranular fracture at a later phase (Fig. 4(d)). The development of intergranular fracture is attributable to the carbide particles on the grain boundary, which reduce the grain boundary strength. The stressstrain hysteretic response obtained in all of the tests in Tables 2 and 3 was essentially elastic, even if the stress amplitude was considerably larger than the monotonic tensile yield stress. No signicant changes of the hysteresis loop were observed during the test. An example is given for the axialtorsional test on specimen A1 in Fig. 5. The linear behavior during unloading and reloading is expected to span over the

range of tensile yield stress plus compressive yield stress in view of the Baushinger eect in elasticplastic materials. It is expected for this material that this range will be extended due to the substantially larger linear response on the compression side. The fatigue parameter versus life curve was determined from the results of uniaxial and torsional fatigue tests shown in Table 2. The fatigue curve is given in Fig. 6 (solid symbols), and fatigue properties are found to be: r0f 4100 MPa, b 0:079, k1 0:77. The fatigue criterion in the presence of mean stress was not possible to be described with a single equation. The following two equations are proposed for tensile and compressive mean stress eects, respectively:   rm 0 n Drn k1 Drh rf 1 k2 0 2Nf b for rm 2 n ! 0; rf   rm 0; 3 Drn k1 Drh r0f 1 k3 n 2Nf b for rm n Drn where rm n is the mean normal stress on the plane of maximum normal stress range, k2 and k3 are material constants. For tensile mean stresses, the parameter versus life relation had the same slope as the no mean stress case. It is noted that Eq. (2) is in the same form as Morrows mean stress model [20] other than the presence of a constant k2. The compressive mean stress data yielded changing slopes in the fatigue curve. Instead of dening the exponent b as a function of Nf, the fatigue strength coecient was set to be dependent not only on the mean stress but also on the normal stress range. This was deduced from the assumption that the tension-

Fig. 5. Stressstrain hysteresis loops in axialtorsional loading test (A1).

Fig. 6. Fatigue parameter versus life relations with and without mean stress.

688

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

undergoing portion in a stress cycle, which can be expressed in terms of the total stress amplitude and mean stress, is responsible for the major fatigue damage. The separation of the carbidematrix interface is considered to be the controlling damage mechanism, and it is most likely to occur under tensile interface stresses. The correlation results are also shown in Fig. 6. The proposed model provides the general trend of mean stress eects. The constants k2 and k3 were found to be 3.49 and 2.49, respectively. It is apparent in Fig. 6 that compressive mean stress is more benecial at high cycles than at low cycles, while tensile mean stress results in the same degree of detrimental eects at both low and high cycles. The considerable scatter found in life data would be related to the statistical nature of carbide size and distribution. More scatter is found with compressive mean stress data, for which the slope of the fatigue curve is small. The results of life prediction are summarized in Table 4 along with the experimental lives. It is found that the angles of fracture surface, measured from the plane normal to the specimen axis, agree reasonably well between predictions and measurements. The scatter in the data considerably obscures the phase-angle dependence of life. However, it seems certain that life is shorter for test conditions closer to in-phase loading (sux 1) than for tests under more out-of-phase loading (suces 2 and 3). The proposed fatigue parameter models this trend correctly. Under strain control tests on ductile metals, it is usually found in the low cycle fatigue regime that out-of-phase loading is more damaging than in-phase loading. This may be attributed to higher stress amplitudes due to additional hardening under nonproportional loading. The stress control
Table 4 Summary of test results and predictions Specimen no. A1 A2 A3 B1 B2 B3 C1 C2 C3 D1 D2 D3 E1 E3-1 E3-2 F3 G1 Life (cycles) Experimental 3044 46,794 45,385 5147 6040 20,609 2139 13,727 6890 12,130 45,731 13,842 20,174 279,968 110,207 8889 124,234 Predicted 1252 5709 9579 1450 14,276 45,150 1383 6281 20,514 3878 36,950 19,832 3315 89,102 87,984 3069 53,996 Fracture angle ( ) Experimental 25 10 172 31 25 143 41 40 135 21 10 170 39 32 35 40 31 Predicted 24 12 174 32 38 140 37 41 137 24 9 168 31 39 39 42 34
v

employed in this study does not dier from strain control since the material response is linear and there is no cycle-dependent change. Thus, the life trend between in-phase and out-of-phase loading is reversed for the present material from that of ductile materials under strain control in the low cycle regime. An interesting comparison of the present results with available data can be seen in high-cycle in-phase and out-of-phase fatigue, where deformation is basically elastic as in the current study. Such data can be found in Refs. [21,22]. The determination of whether in-phase or out-of-phase loading is more damaging at high cycles is dependent on the material. Nishihara and Kawamoto [21] reported that mild steel, hard steel and cast iron exhibited more damage under in-phase loading than out-of-phase loading at 107 cycles under combined bending and torsion. The results of this study are consistent with their results. However, duralumin tested by the same authors [21] did not show phase-angle dependency at 107 cycles, and the high strength steel 42CrMo4V tested by Lempp [22] showed that outof-phase loading is more damaging at 2 106 cycles in bending-torsion tests. The predicted lives were compared with experimental lives in Fig. 7. It is observed that the parameter correlates the data but with some conservatism. The dotted line represents a band of life factor 3. Most of the data points fall within or very close to this band. It is also noted that there is more scatter in life data than usually observed in ductile metals. This is believed to be an inherent character of materials whose life is controlled by defects [7,23].

Fig. 7. Comparison of predicted and experimental fatigue lives under axialtorsional loading.

K.S. Kim et al. / International Journal of Fatigue 26 (2004) 683689

689

6. Conclusions A series of axialtorsional fatigue tests were conducted on 5% chrome steel heat-treated for wear resistance. This material is widely used for work rolls in strip milling operations. The following conclusions can be drawn: 1. The test material fails in a brittle manner with insignicant plasticity being developed under fatigue loading. Fracture occurs on the maximum tensile stress plane. 2. The crack tends to initiate at a cluster of carbide particles on or near the surface. The crack propagates in a transgranular mode initially followed by a mixed transgranularintergranular mode at a later stage. 3. A multiaxial fatigue parameter has been developed in terms of the maximum normal stress range and hydrostatic stress range. This parameter correlates experimental lives with most data points in a band of factor 3. 4. Under axialtorsional fatigue loading with a given ratio of stress amplitudes, in-phase loading produces greater damage than out-of-phase loading for the test material. 5. Compressive mean stresses are more benecial at high cycles than at low cycles, while tensile mean stresses are detrimental to the same extent at low and high cycles. Separate models have been proposed for tensile and compressive mean stress eects.

Acknowledgements This study was supported partly by the Brain Korean 21 Project, and partly by Posco Company. The authors are grateful for the support. The authors also thank Y.C. Park of R&D Center, Doosan Heavy Industries and Construction Company, for his help in preparing the specimens used in the study.

References
[1] Ott GA. The application, metallurgy and maintenance of high hardness, ultra-deep-hardened forged steel work rolls. I&SM 1997;February:2733. [2] Hayashi Y, Hino S, Mizoguchi T, Toyoda H. Crack resistance of deep hardened work rolls for cold strip mills. Kobe Steel Technical Bulletin 1997;33(3):214 [in Japanese]. [3] Ohhashi S, Ishiguro T, Gotoh H. Metallurgical factors aecting thermal shock cracking resistance of work roll for cold strip mills. Iron and Steel 1991;77(5):6529 [in Japanese].

[4] Tallian TE. Simplied contact fatigue life prediction model part I: review of published models. Journal of Tribology 1992;114:20713. [5] Alfredsson B, Olsson M. Applying multiaxial fatigue criteria to standing contact fatigue. International Journal of Fatigue 2001;23:53348. [6] Salehizadeh H, Saka N. The mechanics of crack initiation at hard particles in rolling line contacts. Journal of Tribology 1992;114:3417. [7] Muerling F, Melander A, Tidesten M, Westin L. Inuence of carbide and inclusion contents on the fatigue properties of high speed steels and tool steels. International Journal of Fatigue 2001;23:21524. [8] Hanlon DN, Rainforth WM, Sellars CM. The eect of processing route, composition and hardness on the wear resistance of chromium bearing steels in a rollingsliding conguration. Wear 1997;203204:2209. [9] Haigh BP. The thermodynamic theory of mechanical fatigue and hysteresis in metals. Reports of the British Association for the Advancement of Science, London. 1923, p. 35868. [10] Sines G. Failure of materials under combined repeated stresses with superposed static stresses. Technical note 3495. Washington DC: National Advisory Council for Aeronautics; 1955. [11] Libertiny GZ. Short life fatigue under combined stresses. Journal of Strain Analysis 1967;2(1):915. [12] Davis EA, Connelly FM. Journal of Applied Mechanics 1959;81:25. [13] Manson SS, Halford GR. Journal of Engineering Materials and Technology 1977;99:283. [14] Kalluri S, Bonacuse PJ. In: McDowell DL, Ellis R, editors. Advances in multiaxial fatigue. ASTM STP 1191. 1993, p. 133. [15] Kakuno H, Kawada Y. A new criterion of fatigue strength of a round bar subjected to combined static and repeated bending and torsion. Fatigue and Fracture in Engineering Materials and Structures 1979;2(2):22936. [16] Dang-Van K. Macromicro approach in high-cycle multiaxial fatigue. In: McDowell DL, Ellis R, editors. Advances in multiaxial fatigue. ASTM STP 1191. 1993, p. 12030. [17] Wang CH, Brown MW. A path-independent parameter for fatigue under proportional and nonproportional loading. Fatigue and Fracture in Engineering Materials and Structures 1993;16(12):128598. [18] Socie D. Multiaxial fatigue damage models. Journal of Engineering Materials and Technology, Transactions of the ASME 1987;109:2938. [19] Kim KS, Lee BL, Park JC. Biaxial fatigue of stainless steel 304 under irregular loading. In: Halford GR, Gallagher JP, editors. Fatigue and fracture mechanics. ASTM STP 1380, vol. 31. 2000, p. 7993. [20] Morrow J. Fatigue properties, drafted for SAE publication 339; 1964. [21] Nishihara T, Kawamoto M. The strength of metals under combined alternating bending and torsion with phase dierence. Memoirs of the College of Engineering, Kyoto Imperial University 1945;11(5):85112. [22] Lempp W. Festigkeitsverhalten von stahlen bei mehrachsiger dauerschwingbeanspruchung durch normalspannungen mit ueberlagerten phasengleichen und phasenverschobenen schubspannungen. Dissertation, University, Stuttgart, 1976. [23] Nadot Y, Mendez J, Ranganathan N, Beranger S. Fatigue life assessment of nodular cast iron containing casting defects. Fatigue and Fracture in Engineering Materials and Structures 1999;22:289300.

You might also like