You are on page 1of 9

JFAPBC (2005) 6:30-38 DOI: 10.

1361/154770205X76349 Genesis of Spalling in Tandem

Mill Work-Rolls

(continued)

© ASM International 1547-7029 / $19.00

Genesis of Spalling in Tandem Mill Work-Rolls: Some Observations in Microstructural Degeneration


M.S. Prasad, S.K. Dhua, C.D. Singh, and Amitava Ray
(Submitted August 12, 2005; in revised form October 16, 2005)

Macroscopic investigations of spalled 3% Cr-variety forged steel work-rolls used in the tandem mill of an integrated steel plant showed steel strip welding on roll surfaces. Microstructural observations of roll samples at regions away from strip-welded zones showed the desired uniform dispersion of fine globular carbides in tempered martensite. Quantitative image analysis of all investigated rolls also showed desirable carbide characteristics, with >4 vol.% carbides and >200,000 individual carbides/ mm2. The carbide sizes ranged from 0.69 to 0.83 µm. In contrast, optical and scanning electron microscopy (SEM) observations of the strip-welded regions showed microstructural degeneration thought to have occurred from surface and/or subsurface damage caused by localized thermal shock and intense pressure. This possibly resulted in the formation of a rehardened and heavily retempered zone at the strip-welded region. Cracks originated in the heavily retempered zone due to residual tensile stresses and propagated under the applied rolling stresses to produce spalling. Keywords: microstructural degeneration, spalling, tandem mill, work-rolls

Introduction
Forged hardened and tempered steel work-rolls are generally used in the finishing stands of multistand cold rolling mills for the production of flat products. For effective performance in cold rolling operations, work-rolls should have high surface hardness and an adequate depth of hardening to assure the requisite wear resistance.[1] In addition, work-rolls must have sufficient resistance against rolling-contact fatigue and adequate toughness to prevent premature failures due to spalling.[2] To achieve the desired properties, the manufacture of superior-performance rolls has primarily concentrated on alloy design and heat treatment. The alloy design approach is based on increasing the chromium content of steel to achieve higher hardness, greater hardenability, and increased wear resistance and rolling-contact fatigue strength.[3] Concurrently, heat treatment practices involving induction hardening and tempering have been developed to achieve high surface hardness, a desirable microstructure, and a favorable stress condition on the roll surface. Regardless of heat treatment or alloy content, the microstructure of a hardened and

tempered steel work-roll usually consists of a dispersion of alloy carbides in a matrix of tempered martensite and very little retained austenite.[4] A hardened steel roll, even after tempering, contains high internal residual stress, which, when coupled with the high surface hardness, often renders the roll thermally unstable, crack-sensitive, and vulnerable to spalling.[5] In multistand cold rolling mills, such as the tandem mill, the performance of work-rolls is determined by the metallurgical quality and by mill operational conditions. This is particularly important in present-day cold rolling mills where workrolls are subjected to severe conditions, including high mill speeds, high loads, and stringent requirements for mill productivity and product quality.[6] The severe service conditions imposed on workrolls accentuate the propensity for premature failures, notably by spalling. Spalling is a type of fatigue failure that involves fracturing away of large or small metal pieces from localized areas of the roll working surface.[7] Spalled rolls are generally conditioned by grinding for reuse, although continued spalling may often result in the early

M.S. Prasad, S.K. Dhua, C.D. Singh, and Amitava Ray, R&D Center for Iron & Steel, Steel Authority of India Limited, Doranda, Ranchi-834002, India. Contact e-mail: a123ray@yahoo.com and spmokkapaty@yahoo.co.uk.

30

Volume 5(6) December 2005

Journal of Failure Analysis and Prevention

scrapping of rolls.[8] In the past, the causes of roll spalling have been attributed to mechanical damage, thermal damage, and hydrogen embrittlement; there is, however, no general consensus or agreement on the failure mechanism or on whether the failures initiate at or below the roll working surface.[1,9] Studies have revealed that the main consumption factor for work-rolls in cold rolling mills is not abrasive wear but rather cracks and spalls on the roll surface. Incidentally, 60% of roll consumption has been attributed to accidental damage caused by strip welding or skidding.[10] Moreover, recent investigations of cold mill work-rolls that exhibited varying performances under similar mill operating environments have elucidated the influence of microstructure on spalling propensity and roll life.[11,12] Spalling is a complex phenomenon that may be accentuated by inappropriate mill operating practices and/or poor roll material quality. In order to understand the genesis of work-roll spalling that occurred during a specific rolling campaign, extensive metallurgical investigations were carried out on prematurely failed roll samples. The samples investigated were from 3% Cr-variety forged steel work-rolls (nominal diameter: 585 mm; barrel length: 1420 mm) used in the finishing stands of a five-stand tandem mill in an integrated steel plant under the Steel Authority of India Limited. This paper provides a microstructural insight into the macro- and microstructural manifestations of spalled roll samples and elucidates the possible causes of failure.

microstructural observation in the light microscope, the spalled roll specimens were prepared by conventional metallographic procedures and etched in 5% nital reagent (5 mL concentrated HNO3 in 95 mL ethanol). Microscopic examinations were carried out under bright-field illumination at 500 and 1000× magnifications to observe the characteristics of the phases present. Additionally, microhardness measurements (Vickers pyramid number, or VPN) were made on the cross section of metallographically prepared spalled roll specimens. The microhardness measurements were made using automatic test equipment equipped with a pneumatically actuated diamond indenter; a load of 50 gram-force was applied for the measurements. Scanning electron microscopic observations on polished and nital-etched spalled specimens at different magnifications were used to determine microstructural features at the surface and subsurface zones as well as to observe crack contours. The fracture surfaces of the spalled roll specimens were also observed in the SEM to study topographic features.

Results and Discussions


Roll Life The service life (Table 1) of the six spalled tandem mill work-rolls investigated ranged from 873 to 34,136 t compared to an expected life of 30,000 t. In general, the bulk of the rolls (T2 to T6) exhibited poor performance in the mill, as manifest by the low life values (873 to 9697 t). The highest roll life (34,136 t) was achieved by roll T1 and was attributed to its extensive use prior to the specific campaign period associated with the abnormal incidence of spalling. The scrap diameters
Table 1 Roll Life and Spalling Data of Tandem Mill Work-Roll Samples
Scrap Diameter, mm 564.75 576.60 578.70 580.80 583.00 584.60 Roll Life, t 34,136 9,697 6,004 7,805 2,839 873

Experimental
Spalled pieces were collected from the barrel portion of six tandem mill work-rolls for microstructural and fractographic examination. The particulars of the spalled roll samples (T1 to T6) with respect to their starting diameter, scrap diameter, and roll life (i.e., tonnage rolled in metric tonne, or t) were collected from the roll shop and are shown in Table 1. The spalled samples were ultrasonically cleaned with acetone to remove rolling debris and lubricant prior to examination. Subsequently, specimens were cut from the working surface as well as from the radial cross section for microstructural examination using light and scanning electron microscopy. For

Roll Starting Diameter, Sample mm T1 585.40 T2 T3 T4 T5 T6 586.00 584.60 584.50 586.00 585.60

Journal of Failure Analysis and Prevention

Volume 5(6) December 2005

31

Genesis of Spalling in Tandem Mill Work-Rolls


of rolls T2 to T6 (Table 1) ranged from 576.60 to 584.60 mm; these values were very close to the commissioning diameter of 585 mm and signified low use in the mill prior to spalling-induced degradation. The stipulated scrap diameter (i.e., the diameter corresponding to the desired hardness depth) for such work-rolls in normal use is 545 mm; the scrap diameters of failed rolls (Table 1) are therefore indicative of inadequate use and poor service life owing to spalling-induced failures.

(continued)

Roll Chemistry The chemical compositions (Table 2) of the six spalled roll samples (T1 to T6) showed variation of carbon from 0.82 to 0.92 wt.%, chromium from 2.94 to 3.21 wt.%, molybdenum from 0.19 to 0.24 wt.%, vanadium from 0.12 to 0.16 wt.%, and nickel from 0.11 to 0.22 wt.%. These ranges correspond to the broad compositional range for 3% Cr-base forged steel work-rolls. Increased chromium content in steel work-rolls is known to confer higher wear resistance, ensure greater hardenability, and facilitate the formation of M7C3type carbides. An increase in chromium content from 1.75 to 3.25 wt.% changes the lowerhardness (840 to 1100 VPN) M 3C carbides to higher-hardness (1200 to 1600 VPN) M7C3-type carbides, which are crucial to the wear resistance of steels.[6] The presence of nickel in the work-roll composition confers hardenability; its content, however, is restricted in steels because nickel promotes the formation of retained austenite, which reduces hardness and is undesirable from the standpoint of roll performance. The presence of molybdenum in the steel composition (0.19 to 0.24 wt.%) is primarily aimed at improving the hardenability, temper resistance, and wear resistance of the tandem mill work-rolls.
Table 2

Macrostructural Features The visual appearance of the working and fracture surfaces of all six investigated spalled roll samples was similar; the working surfaces of these roll samples exhibited metal sticking, indicative of strip welding during rolling. Figures 1(a) and (b) show the working and fracture surfaces of a typical spalled sample and illustrate the presence of strip welding (Fig. 1a) on the working surface of the roll. Furthermore, the roll working surface showed mechanical damage marks that were probably

Fig. 1

Spalled roll sample T4 showing (a) working surface and (b) fracture surface

Chemical Composition of Spalled Tandem Mill Work-Roll Samples


C 0.82 0.88 0.91 0.92 0.86 0.87 Mn 0.34 0.29 0.28 0.26 0.27 0.44 Si 0.44 0.37 0.43 0.32 0.36 0.54 P 0.020 0.015 0.013 0.015 0.012 0.013 Composition, wt.% S Cr 0.015 0.011 0.010 0.019 0.010 0.0111 2.94 3.21 3.10 2.99 3.12 3.05 Mo 0.21 0.20 0.19 0.23 0.24 0.22 Ni 0.21 0.18 0.22 0.14 0.19 0.11 Al 0.016 0.011 0.01 0.006 0.013 0.006 V 0.13 0.15 0.14 0.16 0.13 0.12

Roll Sample T1 T2 T3 T4 T5 T6

32

Volume 5(6) December 2005

Journal of Failure Analysis and Prevention

generated as a result of skidding. It is apparent from Fig. 1(b) that the fracture surface shows fatiguelike characteristics followed by final overload fracture. The macroscopic evidence of steel strip welding to the working surface of the roll (Fig. 1a) and the fracture pattern on the spalled surface (Fig. 1b) are possibly indicative of a tandem mill wreck. The interaction of a propagating subsurface fatigue crack with the wreck generated a shock wave. The micrographs in Fig. 2 to 4 demonstrate the presence of subsurface cracks, and such crack formation indicates the presence of a microstructural region that is susceptible to cracking. In many of the spalled roll samples, it was difficult to determine the crack origin; however, the one spalled roll piece (Fig. 1a and b) demonstrates a case in which the origin of the crack was near the roll surface. After initiation, the crack propagated into the roll at an oblique angle and then continued to run almost parallel to the roll surface. This crack pattern is typical of rolling-contact fatigue.

count (201,342 to 268,110 number/mm 2) and size (0.69 to 0.83 µm) parameters were found to be in close range and were typical of good-quality forged steel rolls.[11,12] Because the microstructure and quantitative attributes of microstructural constituents in all the investigated roll samples (at regions away from strip-welded zones) were similar

Fig. 2

Microstructural Features Figure 5 shows the nital-etched microstructure of a typical spalled tandem mill work-roll (sample T4) away from the strip-welded region. The microstructure in this region shows a uniform dispersion of fine globular carbides in a matrix of tempered martensite. This type of microstructure is desirable for goodquality cold mill work-rolls.[13] In fact, the microstructures of the working surface of all the investigated spalled rolls were similar to that observed in sample T4. The quantitative image analysis data pertaining to carbide content (volume fraction), carbide count (number/mm2), and carbide size (µm) of the investigated roll samples and pertaining to regions away from the strip-welded spot are furnished in Table 3. It is evident from this table that carbide content in all the spalled roll samples was greater than Fig. 3 Scanning electron micrographs of spalled roll sample T4 showing (a) structure 4 vol.% and ranged from 4.32 to at some other location, (b) rehardened zone, (c) less rehardened zone, (d) heavily tempered zone, and (e) low-temperature tempered zone 5.42%. Furthermore, the carbide

Optical micrograph of spalled roll sample T4 at stripwelded region showing different microstructural features. 500×

Journal of Failure Analysis and Prevention

Volume 5(6) December 2005

33

Genesis of Spalling in Tandem Mill Work-Rolls

(continued)

Fig. 4

SEM images showing crack initiation, propagation, and branching. (a) Region with different structural features. 500×. (b) Montage of crack initiation, propagation, and branching. 2000×

Table 3

Quantitative Image Analysis Data of Spalled Tandem Mill Work-Roll Samples


Carbide count, number/mm2 202,468 227,471 268,110 201,342 244,402 223,058 Carbide size, µm 0.83 0.69 0.71 0.77 0.76 0.82

Roll Carbide content, Sample vol.% T1 T2 T3 T4 T5 T6 4.96 4.95 5.42 4.32 5.35 5.42

Fig. 5

Optical micrograph of nital-etched spalled roll sample T4 away from strip-welded region showing uniform dispersion of fine globular carbides in the matrix of tempered martensite. 1000×

to each other and normal, the possibility that the roll material was defective was ruled out in this case. The optical microstructure (Fig. 2) as well as the SEM micrographs (Fig. 6) of the same spalled roll (sample T4), but at the strip-welded region, revealed microstructural heterogeneity and the presence of different structural features. The

martensitic needles in the strip-welded region demonstrate that this area experienced rehardening and produced a brittle zone at the near-surface region (Fig. 6c). In the optical microscopic image, this zone etches white (Fig. 2) and overlies a less rehardened zone having microstructural features (Fig. 6d) similar to that of the rehardened zone, except for the presence of primary carbides. The SEM photographs (Fig. 6c and d) show that both

34

Volume 5(6) December 2005

Journal of Failure Analysis and Prevention

primary and spheroidized carbides at the stripwelded region are few in number. Directly below the less rehardened zone is a region of heavily tempered martensite with bimodal distribution of carbides (Fig. 6e). This bimodal distribution depicts the presence of very fine carbides (M3C) at needle boundaries, as well as spheroidized carbides in the matrix. This region appears black in the optical microscopic image (Fig. 2). Furthermore, there is the presence of a region having microstructural features comprising low-temperature (<200 °C) tempered martensite along with spheroidized carbides. This region represents the normal roll matrix structure, as shown in Fig. 5 and 6(f ) at 1000 and 4500× magnifications, respectively. In addition to the previously mentioned regions, there is a welded strip zone showing the deformed ferrite-pearlite microstructure (Fig. 6b) characteristic of normal cold-rolled steel. The optical micrograph (Fig. 2) also reveals a crack as well as evidence of brittle fracture emanating from the heavily tempered region at an angle of 45° toward the working surface (outer layers) and the inside layers of the roll, respectively. The SEM images of these microstructures in the radial section of the same spalled roll sample at another location are shown in Fig. 3(a) at 750× magnification. The microstructural features of the rehardened, less rehardened, heavily tempered, and low-temperature tempered regions are shown in Fig. 3(b-e), respectively, at 3000× magnification. This region also contains a crack along the interface between the rehardened and less rehardened zones. Further, Fig. 3(a) and (b) suggest that the crack emanated from the heavily retempered region and moved both outward and inward of the roll cross section. The crack appears to originate at an angle of 45° and changes Fig. 6 path during propagation. The presence of the secondary cracks branching from the main crack in

the rehardened zone as well as along the interface between the rehardened and less rehardened zones was also apparent. Optical (Fig. 2) and scanning (Fig. 3, 4, and 6) microscopic examinations of the spalled roll cross section further indicate damage to the working surface of the roll. The damage to the roll surfaces, in a sense, creates subsurface regions having altered microstructures and properties. These subsurface regions include a rehardened zone near the surface, a less rehardened zone next to the surface layer, a heavily retempered zone directly below the less rehardened zone, and finally, a low-temperature tempered zone representing the normal roll structure. In fact, the rehardened and less rehardened zones appear to have formed as a result of localized reaustenitizing above Accm temperature and quenching to form an untempered martensite. Investigations have proven that the temperature of the roll surface within a depth of 0.2 mm may

Scanning electron micrographs of polished and etched section of spalled roll sample T4 showing (a) surface structure with welded metal strip, (b) welded strip of deformed ferrite-pearlite zone, (c) rehardened zone with freshly formed martensitic structure, (d) less rehardened zone, (e) heavily retempered martensitic zone, and (f ) low-temperature tempered zone

Journal of Failure Analysis and Prevention

Volume 5(6) December 2005

35

Genesis of Spalling in Tandem Mill Work-Rolls


increase to 900 °C. This inevitably results in structural transformations in the thermally influenced zone and produces internal radial and tangential stresses in the roll. These stresses are superimposed on the thermal and shearing stresses in the roll gap, and the total stresses may attain values resulting in cracking and spalling.[10] Figures 1(a), 2, and 6(a) illustrate a situation where steel strip had welded to the roll surface. In consequence, the roll material in contact with the steel strip was reaustenitized and quenched and therefore became harder than the matrix. However, the temperature below the roll surface and subsurface was not sufficient to reaustenitize the material but was higher than the original roll-tempering temperature (rolls are usually tempered in the range of 100 to 260 °C). This region became heavily tempered because the temperature range was approximately 500 to 600 °C. This tempering softened (633 VPN) the material. The degradation of the original roll microstructure in the surface and subsurface regions is also corroborated by the microhardness distribution (Fig. 7) and is consistent with the occurrence of surface damage.

(continued)

This surface damage appears to have originated from localized heating, which seems to be the result of pinching or skidding of the steel strip between the work-rolls or, in more acute cases of a mill wreck, where the strip breaks and wraps around the roll. This pinching, skidding, or mill wreck created a localized thermal shock and intense pressure, which caused the reaustenitizing and quenching of the roll material as well as the welding of the steel strip to the roll surface.[1] The hardness changes (Fig. 7) and the associated residual tensile stresses accompanying localized thermal shock are sufficient to induce cracks in the heavily retempered subsurface region of the roll, and these cracks can then propagate both outward and inward (Fig. 4) to induce spalling.

Fractographic Features The SEM fractographs of regions pertaining to coarse beach, fine beach, and radial marks are shown in Fig. 8(a-c), respectively, at 2000× magnification. The SEM fractograph in Fig. 8(a) is consistent with a brittle fracture. The fracture surface in Fig. 8(b) has features similar to that in Fig. 8(a) but on a smaller scale and is consistent with rolling-contact

Fig. 7

Distribution of microhardness across roll depth

36

Volume 5(6) December 2005

Journal of Failure Analysis and Prevention

fatigue or simple surface damage during the fracture process. In contrast, evidence from intergranular cracking and overload fracture is shown in Fig. 8(c).

Microhardness Figure 7 shows the variation in cross-sectional hardness of a typical spalled roll specimen. It is seen that a rehardened zone (970 VPN) exists very close to the roll surface, along with a comparatively

less rehardened zone (900 VPN) directly under it. Furthermore, the sharp decrease in hardness (633 VPN) marks the presence of a heavily tempered zone just below the less rehardened zone. Finally, the hardness increases to a saturation level (837 VPN), which appears to be the hardness of the low-temperature tempered region. Interestingly, the hardness of the welded cold-rolled steel strip is approximately 425 VPN.

Conclusions
• Macroscopic observations of all investigated work-rolls showed steel strip welding and evidence of mechanical damage on the working surfaces, suggesting mill abuse and/or inappropriate operational conditions. • Microstructural observations of all roll samples at regions away from strip-welded zones showed uniform dispersion of fine globular carbides in tempered martensite and signified a desirable microstructure for such quality rolls. This observation is consistent with the initial use of properly heat treated roll material. • Quantitative image analysis of all investigated rolls at regions away from strip-welded zones showed desirable carbide characteristics, with contents >4 vol.%, counts >200,000 number/ mm2, and sizes ranging from 0.69 to 0.83 µm. This analysis corroborated good metallurgical quality of the rolls. • Optical and SEM observations of strip-welded regions, however, showed microstructural degeneration that possibly occurred from surface and/ or subsurface damage caused by localized thermal shock as well as localized intense pressure that would be associated with pinching, skidding, or a mill wreck. This abusive use of the mill resulted in the formation of a rehardened, a less rehardened, and a heavily retempered zone at the strip-welded region. The cracks originated in the heavily retempered zone due to a combination of residual and applied stresses and propagated under applied rolling stresses to cause premature failure by spalling.

Fig. 8

Scanning electron micrographs of spalled roll sample T4 showing (a) brittle fracture surface, (b) brittle zone with very small cleavage facets, and (c) overload fracture region. 2000×

Acknowledgments
The authors are grateful to the management of

Journal of Failure Analysis and Prevention

Volume 5(6) December 2005

37

Genesis of Spalling in Tandem Mill Work-Rolls


the R&D Center for Iron and Steel (RDCIS), Steel Authority of India Limited, Ranchi, for the encouragement and support in pursuing this study. Thanks are expressed to the personnel of the metallography laboratory, notably U.N. Jha, C.B. Sharma, and J. Guria, at RDCIS for their help in the investigation and to B. Khalkho for typing the manuscript in a short span of time.

(continued) 6. G.A. Ott: “The Physical Metallurgy of 4% Chromium Forged Steel Cold Mill Work Rolls,” 43rd Mechanical Working and Steel Processing Conf. Proc., vol. 39, Iron and Steel Society, Warrendale, PA, 2001, pp. 747-80.

7.

References
1. J.M. Chilton and M.J. Roberts: “Factors Influencing the Performance of Forged Hardened Steel Rolls,” Iron Steel Eng., Jan 1981, pp. 77-82. K.F. Reppert and B. Somers: “Development of a Deep Hardening Work Roll at Lehigh Heavy Forge,” 43rd Mechanical Working and Steel Processing Conf. Proc., vol. 39, Iron and Steel Society, Warrendale, PA, 2001, pp. 731-45. S. Kawashima, M. Yoshikawa, and S. Izumikawa: “New Trend of Forged Hardened Steel Work Rolls and Back Up Rolls for Rolling Mills in Japan,” 29th Mechanical Working and Steel Processing Conf. Proc., vol. 25, Iron and Steel Society, Warrendale, PA, 1988, pp. 49-57. R.L. Bodnar, M. Lin, and S.S. Hansen: “The Physical Metallurgy of Forged Cold-Mill Work-Roll Steels,” 33rd Mechanical Working and Steel Processing Conf. Proc., vol. 29, Iron and Steel Society, Warrendale, PA, 1992, pp. 171-85. M. Nakagawa, A. Hoshi, A. Asano, and Y. Nambu: “Causes and Countermeasures of Spalling of Cold Mill Work Rolls,” Iron Steel Eng., Mar 1981, pp. 44-49.

2.

S.J. Manganello and D.R. Churba: “Roll Failures and What to Do When They Occur,” Iron Steel Maker, Dec 1980, 7(12), pp. 26-34. 8. D. Mukherjee, A. Ray, and S.K. Bhattacharyya: “Spalling Resistance of Forged Steel Cold Rolling Mill Rolls: A Microstructural Insight,” Mater. Forum, Dec 1992, 16(4), pp. 317-25. 9. A.N. Sinha and C.S. Sivaramakrishnan: “Failure of Work Roll of Cold Rolling Mill,” Ironmaking Steelmaking, 1994, 21(2), pp. 154-57. 10. W. Patt: “Production and Properties of Hi-Cr Rolls and Their Behavior in Cold Strip Mills,” 26th Mechanical Working and Steel Processing Conf. Proc., vol. 22, Iron and Steel Society, Warrendale, PA, 1985, pp. 101-07. 11. M.S. Prasad, A. Ray, S.K. Dhua, R. Avtar, and S. Jha: “Premature Failure of Work-Rolls in Tandem Mill: Some Microstructural Revelations,” J. Fail. Anal. Prevent., Jun 2004, 4(3), pp. 67-72. 12. A. Ray, M.S. Prasad, P.K. Barhai, and S.K. Mukherjee: “Microstructural Characteristics of Prematurely Failed Cold-Strip Mill Work-Rolls: Some Observations on Spalling Susceptibility,” J. Mater. Eng. Perform., Apr 2005, 14(2), pp. 194-202. 13. G.A. Ott: “The Development of Forged Hardened Steel Roll Metallurgy to Meet Special Rolling Mill Requirements,” 33rd Mechanical Working and Steel Processing Conf. Proc., vol. 29, Iron and Steel Society, Warrendale, PA, 1992, pp. 159-70.

3.

4.

5.

38

Volume 5(6) December 2005

Journal of Failure Analysis and Prevention

You might also like