Vous êtes sur la page 1sur 14

This article was published in an Elsevier journal.

The attached copy is furnished to the author for non-commercial research and education use, including for instruction at the authors institution, sharing with colleagues and providing to institution administration. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elseviers archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Author's personal copy


ARTICLE IN PRESS

International Journal of Machine Tools & Manufacture 48 (2008) 402414 www.elsevier.com/locate/ijmactool

Modeling of white layer formation under thermally dominant conditions in orthogonal machining of hardened AISI 52100 steel
Anand Ramesh, Shreyes N. Melkote
The George W. Woodruff School of Mechanical Engineering, Georgia Institute of Technology, Atlanta, GA, USA Received 20 December 2006; received in revised form 15 September 2007; accepted 24 September 2007 Available online 6 October 2007

Abstract This paper presents a nite element model for white layer formation in orthogonal machining of hardened AISI 52100 steel under thermally dominant cutting conditions that promote martensitic phase transformations. The model explicitly accounts for the effects of stress and strain, transformation plasticity and the effect of volume expansion accompanying phase transformation on the transformation temperature. Model predictions of white layer depth are found to be in agreement with experimental values. The paper also analyzes the effect of white layer formation on residual stress evolution in orthogonal cutting of AISI 52100 hardened steel. Model simulations show that white layer formation does have a signicant impact on the magnitude of surface residual stress and on the location of the peak compressive residual stress. r 2007 Elsevier Ltd. All rights reserved.
Keywords: Finite element modeling; White layer formation; Phase transformation; Hard machining; Orthogonal cutting

1. Introduction Considerable attention has been focused in recent years on understanding microstructure changes observed in the surface layers of machined hardened steels. The microstructure change is commonly referred to as white layer or white etching layer since it appears white when observed in an optical microscope after standard metallographic preparation [1]. The layer is typically a few microns thick and is hard and brittle [2,3]. Due to common belief that this layer is detrimental to fatigue life of the part, current industry practice is to supernish the machined surface to remove the white layer. Alternatively, it is desirable to develop a model capable of predicting the onset of white layer formation as a function of the machining conditions. Such a model can then be used to identify cutting conditions that will not result in white layer or will produce minimal white layer. Over the years, a number of experimental investigations aimed at understanding the formation mechanisms and
Corresponding author.

E-mail address: shreyes.melkote@me.gatech.edu (S.N. Melkote). 0890-6955/$ - see front matter r 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.ijmachtools.2007.09.007

properties of white layer in machining and grinding have been reported [417]. However, very limited work has been reported on modeling of white layer formation in machining of hardened steels. Modeling of hard machining is incomplete without white layer formation because of its possible effect on the evolution of residual stresses in the machined part surface. Although work on modeling of the hard turning process to predict forces [18], chip formation [19,20] and tool wear [21] has been reported, only a handful of researchers have attempted to explicitly model white layer formation in machining and/or grinding of hardened steels. Mahdi and Zhang [2224] utilized a nite element (FE) framework to predict phase transformations in grinding. Akcan [25] and Chou and Evans [26] used an analytical approach to predict white layer formation by solving a moving heat source problem. These efforts assumed that white layer formation is due to thermally driven phase transformation effects. However, they did not consider the mechanical effects of stress and strain on transformation temperatures, effect of volume expansion and transformation plasticity associated with martensitic transformation and the impact of white layer formation on residual stresses.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 403

The importance of including these effects in the simulation of hard machining is discussed below. Phase transformation in steel is accompanied by effects such as transformation plasticity and volume expansion. Transformation plasticity is an apparent increase in ductility accompanying austenitemartensite transformation in steels, while the volume expansion corresponds to a 4.4% volume increase associated with martensite formation [27,28]. This apparent increase in ductility has been explained as a pseudo-plasticity effect that arises from the change in material yield stress over the transformation temperature range [29]. The improvement in ductility due to transformation plasticity is important in machining since it translates to increased strain, which in turn affects the ow stress. Improved ductility at high deformation rates also renes the grain size, further modifying the material properties. These effects are routinely incorporated in heat treatment simulations of steels [2934] but have not been included in models of white layer formation in machining. It is known that defects such as dislocations act as nucleation sites for martensite. This suggests that deformation of the material is benecial to martensite transformation [27]. Externally applied stress also increases the number of nucleation sites due to dislocation formation thereby assisting phase transformation [35]. Specically, tensile deformation increases the martensite start (Ms) temperature, while shear deformation assists transformation, martensite formation itself being a shear process [36]. Such effects have not been considered in models of white layer formation in machining of hardened steels. Another factor that has not been considered is the close relationship that exists between phase transformations and the resulting stressstrain evolution in the machined surface. Therefore, simulations of residual stress in the workpiece are incomplete without the effects of white layer formation. This paper presents an FE model of white layer formation in orthogonal machining of AISI 52100 hardened steel that incorporates the effects of stress and strain on transformation temperature, volume expansion and transformation plasticity. The model is valid for continuous white layer formation under thermally dominant cutting conditions where phase transformation is likely to occur. Comparisons of model predictions with experimental data are presented and the results discussed. 2. Development of thermo-mechanical FE model The white layer model development in this paper employs an FE-based thermo-mechanical model formulation of the orthogonal machining process. The explicit dynamics FE procedure is used to develop the thermo-mechanical model of orthogonal cutting. ABAQUSs/Explicit v6.2 is used for all FE simulations reported in this paper. Plane strain four-node quadrilateral elements of type CPE4RT are used to model the workpiece (AISI 52100, 62 HRC) and the tool (cBN). The workpiece

is modeled as elasticplastic, whereas the tool is modeled as purely elastic. The ow stress characteristics of the workpiece material are modeled by the JohnsonCook equation [37]. The coefcients of the ow stress equation were developed using the procedure outlined by Shatla et al. [38]. The nominal chemical composition and physical properties of the workpiece and tool materials used are given in Tables 13, respectively. In order to aid the initial progress of the simulation, the orientations of elements in the chip prior to the start of machining are altered as shown in Fig. 1 through the use of Bezier curves [39]. As can be seen, the basic shape of the chip is not changed. Only the conguration of the elements within the chip is altered. The boundary conditions imposed on the model are shown in Fig. 2. The nodes along the left (AB), bottom (BC) and right side (CD) of the workpiece are constrained against movement in the y-direction. They are given a
Table 1 Nominal chemical composition of AISI 52100 steel (Metals Handbook, 1990) Element % Vol. C 0.981.1 Cr 1.4 Fe 97.05 Mn 0.35 Si 0.25 P o0.25 S o0.25

Table 2 Physical properties of workpiece material (AISI 52100, 62 HRC) Density (kg/m3) Specic heat (J/kg 1C) 7827 458, 25oTo204 640, 204oTo426 745, 426oTo537 798, T4537 46.6 11.5, 12.6, 13.7, 14.9, 15.3, 210 0.277 688.17 150.82 0.3362 0.04279 2.7786 25oTo204 204oTo398 698oTo704 704oTo804 T4804

Thermal conductivity (W/m K) Coefcient of thermal expansion (/1C, 106)

Youngs modulus (GPa) Poissons ratio JohnsonCook constants A (Mpa) B (Mpa) n C m

Table 3 Physical properties of Kennametal KD120 grade cBN Density (kg/m3) Specic Heat (J/kg 1C) Thermal conductivity (W/m K) Coefcient of thermal expansion (/1C, 106) Youngs modulus (GPa) Poissons ratio 3420 750 100 4.9 680 0.22

Author's personal copy


ARTICLE IN PRESS
404 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414

velocity in the x-direction equal to the cutting speed. The nodes on the top and right side of the tool, denoted by MN and NO, are constrained in all directions. An initial temperature of 25 1C is applied to all nodes in the model. Automatic time increment is used for the solution procedure. The model is solved via a fully coupled thermo-mechanical simulation. The effective plastic strain criterion was used for the simulation of chip formation. A baseline value is obtained from the formula proposed by Guo and Dornfeld [40]: 2 cos ae pl p , f 31 sin ae (1)

where ae is the rake angle of the tool. For the cutting geometry used in this study, a value of 1.15 is obtained as a baseline value for the effective plastic strain for chip formation. In reality, the effective plastic strain at failure varies with strain, strain rate and temperature, all of which vary with cutting conditions. Therefore, the sensitivity of model prediction to the value of the effective plastic strain was evaluated via model simulations carried out at a cutting speed of 122 m/min,

undeformed chip thickness (feed) of 0.127 mm, 1 mm width of cut and 01 rake angle. In these simulations, the behavior of the FE mesh around the tool edge and the force predictions was examined by varying the effective plastic strain for failure over a range of values [41]. From this analysis, a value of 1.15 was found to work well for feeds less than 0.152 mm while a value of 1.25 was found to work well for feeds greater 0.152 mm. This is attributed to the variation of pl with strain, strain rate and temperature, as f would occur with a change in feed. Note that Eq. (1) does not capture this probable variation. All toolworkpiece frictional interactions are modeled using the Coulomb friction model, which is stated mathematically as t ms t tcrit when msotcrit sliding, when ms tcrit sticking, (2)

where t is the frictional shear stress, m is the coefcient of friction, s is the frictional normal stress and tcrit is the shear ow stress of the workpiece material and is equal to p (3) tcrit sy = 3, where sy is the uniaxial ow stress of the material. As shown elsewhere [41], suitable values for tcrit were obtained from Oxleys predictive machining theory. Sensitivity studies were performed on this parameter and the coefcient of friction (m). It is found that using the values predicted by Oxleys method for tcrit and using m 0.35 yield the best results. It is assumed that 90% of the energy expended during plastic deformation is converted into heat. Also, frictional heating is accounted for, with all the frictional energy dissipated as heat. A mesh convergence study was also performed. It was found that a coarse mesh (element length 25 mm) sufced for force predictions, but temperatures and plastic strains below the workpiece surface converged to higher values for an element size of 3.125 mm. Validation of the ow stress equation was therefore carried out using the 25 mm mesh,

Fig. 1. Modied region of chip to aid simulation progress.

Fig. 2. Boundary conditions in FE model.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 405

while modeling of white layer formation and residual stresses was carried out using the 3.125 mm mesh.

3. Thermo-mechanical model validation


Fig. 3. Schematic diagram of orthogonal cutting tests carried out on a two-axis CNC lathe.

Table 4 Simulation conditions for thermo-mechanical FE model validation Test no. 1 2 3 4 Speed (m/min) 121.92 152.4 152.4 152.4 Feed (mm) 0.152 0.127 0.152 0.178 tcrit (Mpa) 578.46 580.87 566.34 554.37 ef 1.15 1.15 1.15 1.25

Experimental validation of the thermo-mechanical FE model was carried out at this stage. Dry orthogonal cutting tests were performed on a 25 mm long and 0.9 mm thick tube of hardened 52100 steel (62 HRC) on a two-axis CNC lathe (Hardinge T42SP). A schematic of the test conguration is shown in Fig. 3. As seen in the gure, the tool is fed axially into the tube with the cutting edge of the tool kept perpendicular to the feed and velocity vectors. HighcBN-content inserts (Kennametal NG3125R KD120 grade) with nominally up-sharp edges were used with a 01 rake angle tool holder (Kennametal NER 123B NF1). The length of each cut was xed at 6.35 mm.

Cutting Forces 600 Fc, Experiment 500 Cutting Force, N 400 300 200 100 0 121.92 m/min, 152.4 m/min, 152.4 m/min, 152.4 m/min, 0.152 mm/rev 0.127 mm/rev 0.152 mm/rev 0.178 mm/rev Machining Conditions: Cutting Speed/Feed Thrust Forces 350 Ft, Experiment 300 Thrust Force, N 250 200 150 100 50 0 121.92 m/min, 152.4 m/min, 152.4 m/min, 152.4 m/min, 0.152 mm/rev 0.127 mm/rev 0.152 mm/rev 0.178 mm/rev Machining Conditions: Cutting Speed/Feed
Fig. 4. Comparison of predicted and experimental forces. All forces are in N per mm width of cut. Error bars denote the range of measured forces.

Fc, Oxley Fc, FEM

Ft, Oxley Ft, FEM

Author's personal copy


ARTICLE IN PRESS
406 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414

The cutting conditions used in the tests are shown in Table 4. The same conditions are used in the model simulations along with the values of tcrit and failure strain ef. The value of coefcient of friction used for all conditions listed in the table was 0.35. A 1 mm width of cut was simulated in all cases. Each cutting test was repeated three times and the cutting and thrust forces were measured by a platform-type piezoelectric force dynamometer (Kistler 9257B). A comparison of experimentally determined forces and predicted values is shown in Fig. 4. It can be seen that satisfactory correlation is obtained between the experimental and predicted values. The maximum prediction error in the cutting force is under 3% while the maximum

prediction error in the thrust force is about 13%. The deformed mesh conguration and temperature contours at the end of a typical simulation are shown in Fig. 5. Based on the above results, the thermo-mechanical FE procedure is deemed to be satisfactory for incorporating the phase transformation model of white layer formation, which is discussed next.

4. Modeling of white layer formation Since the formation of white layers in hard machining under thermally dominant machining conditions conducive to the occurrence of phase transformation is to be carried

Fig. 5. Temperature contours and deformed conguration for 152.4 m/min machining speed, 0.178 mm feed validation case.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 407

out, the problem is likened to a quenching process. With this in mind, the methodology to model quenching consists of two parts: (a) Thermal analysis of the quenching process, which is coupled to phase transformation. (b) Mechanical analysis based on the thermal loads generated in the thermal analysis. This includes interaction between the stresses and strains and the progress of phase transformation. Fig. 6 explains the modeling approach. The core of the model is the simulation of thermo-elasto-plastic material behavior. Phase transformation effects are accounted for simultaneously in the simulation. In this manner, full coupling between phase transformation effects and thermo-elasto-plastic material behavior is obtained. The phase transformation model is implemented as a user-dened subroutine (VUMAT) written in Compaq Digital FORTRANs [42] and linked to ABAQUSs v6.2. All calculations are performed using the explicit dynamics procedure with full thermo-mechanical coupling. 4.1. Effect of stress and strain on phase transformation temperatures As noted earlier, the effect of stress and strain on the temperature at which austenite formation starts (As temperature) needs to be accounted for. Grifths [7] proposed the use of the ClausiusClayperon equation to account for the effect of pressure (stress) on the As temperature. The ClausiusClayperon equation describes the effect of pressure on the equilibrium between phases in solid, liquid or gaseous form. For the ferrite/martensite to austenite (a0 g) transformation the equation takes the following form [43]: dP DH tr , dT TDV tr
START

Phase transformation effects

Thermo-mechanical behavior

No Simulation complete?

Yes

STOP
Fig. 6. Layout of model to simulate white layer formation in machining.

where DHtr is the heat of transformation involved in the a0 g transformation, DVtr is the volume change per mole due to transformation and T is the transformation temperature. An approximate but simple analysis based on the above equation can be used to illustrate the impact of pressure (stress) on the ferrite/martensite to austenite transformation temperature. For simplicity, consider that the ag transformation of 52100 steel is similar to that of pure iron (a phase) and neglect the change in DHtr with temperature. Noting that the nominal transformation temperature (T) of pure iron is 910 1C and the corresponding DHtr is 215 cal/g atom [43], the change in volume per mole (DVtr) of the ag transformation can be estimated from the molar weight of pure iron (55.85 g) and the densities of ferrite (7.571 g/cm3) and austenite (7.633 g/cm3) phases to be DVtr 55:8555:85 0.06 cm3. Substituting this and the 7:633 7:571 other quantities into the ClausiusClayperon equation and noting that 1 cal 41.3 cm3 atm yields dT/dP 0.0081/atm [43]. Using a value of 1300 MPa (or 12,833 atm) for the equivalent stress in the shear plane in orthogonal cutting of 52100 steel (derived from FE simulations of orthogonal cutting of 52100 steel) as a measure of the pressure acting on the material in machining, the drop in phase transformation temperature is estimated to be 103 1C. In addition, Beswick [44] has reported a 15 1C drop in the transformation temperature with 64% plastic strain in the material. Since plastic strains in machining can be signicantly higher, a greater reduction in the transformation temperature is likely. Nevertheless, taking 15 1C as a conservative estimate of the reduction in transformation temperature due to plastic strain, the total estimated reduction in the transformation temperature due to pressure (stress) and plastic strain is 118 1C. Using this as a rough estimate of the expected reduction in transformation temperature for AISI 52100 steel yields a nal a0 g transformation temperature of 614 1C (note: the nominal austenitization temperature of 52100 steel is 732 1C). Although approximate, the foregoing analysis clearly indicates that the effect of mechanical stress (pressure) and plastic deformation on the As temperature can be signicant and must be considered in white layer prediction. In order to evaluate the As temperature more accurately in this work, FE simulations were run at 243.84 m/min (800 SFPM) cutting speed with 0.127 and 0.178 mm/rev (0.005 and 0.007 ipr) feeds. It needs to be emphasized that, at these conditions, thermal effects, and hence phase transformations, tend to dominate white layer formation [17]. Temperature proles below the workpiece surface were extracted from the model simulations and temperatures corresponding to the experimentally determined white layer depths were taken as the As temperatures for the two cases. Fig. 7 shows the predicted temperature proles for the two feeds used in the simulations. The vertical lines in the gure represent the measured white layer depths corresponding to the two feeds. This procedure yielded 570.77 and 547.79 1C for the As

Author's personal copy


ARTICLE IN PRESS
408 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414

where F new and F old are the martensite fractions at the end m m and beginning of the increment, respectively. As discussed earlier, an increase in ductility of the material due to transformation plasticity is observed. Transformation plasticity can be modeled as an additional source of strain in the material, whose evolution is described as follows [29]: d TP 2K TP s1 F m dF m ,  (8)

where d TP is the effective transformation plasticity strain,  s is the effective stress and KTP is a material constant equal to 5.08e5. The component-wise transformation plasticity strains may then be calculated as [29]   3 dTP  TP dij (9) Sij , 2 s
Fig. 7. Simulated temperature proles below workpiece surface for cutting speed of 243.84 m/min.

where Sij are the components of the stress tensor. Martensite formation also includes a 4.4% increase in volume. This may be taken into account as follows: dTF 10:044 dF m dij , ij 3
Table 5 Conditions used for prediction of white layer formation Speed (SFPM/m/min) 700/213.36 900/274.32 700/213.36 900/274.32 Feed (ipr/mm/rev) 0.005/0.127 0.005/0.127 0.007/0.178 0.007/0.178

temperatures for the two cases. These values deviate from the value (614 1C) estimated from the foregoing simple analysis by 7% and 10.7%, respectively. The nominal temperature at which martensite formation starts is taken to be 131.44 1C [28]. However, the nominal Ms temperature is affected by stress, which can be modeled as [29] DM s Askk Bs, (4)

(10)

where DMs reects the change in the Ms temperature and A ( 0.05) and B ( 0.033) are material constants. If the temperature of a material point being analyzed exceeds the As temperature, it is assumed to be instantaneously austenitized. The temperatures of all such transformed material points in the FE model are subsequently monitored to check for the possibility of martensite formation. If the temperature falls below the corrected Ms temperature, the fraction of martensite formed is determined from the following equation proposed by Koistinen and Marburger [45]: F m 1 explM s T, (5)

Table 6 Material properties of low-cBN-content tool Density (kg/m3) Specic heat (J/kg 1C) Thermal conductivity (W/m K) Coefcient of thermal expansion (/1C, 106) Youngs modulus (GPa) Poissons ratio 4370 750 4.4 4.9 600 0.17

where Fm is the fraction of martensite formed, l is a material constant equal to 0.011 for most steels and T is the instantaneous temperature. Since the FE procedure is based on an incremental method, the change in the amount of martensite formed due to a drop in temperature dT is given as dF m l explM s T dT. (6)

The quantity of martensite present at any given increment, provided the temperature is below Ms and if austenitization has taken place, is therefore F new F old dF m , m m (7)
Fig. 8. Regions showing phase transformation during machining.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 409

where dTF represents the volumetric strain due to ij martensite formation and dij is the Kronecker delta. In order to incorporate the phase transformation procedure in the elasto-plastic material model developed earlier in the paper, the transformation plasticity strains

and the transformation strains described by Eqs. (9) and (10) need to be included in the strain summation equation for the material as follows: el elold dij dpl dTP dTF . ij ij ij ij ij (11)

White layer White layer Dark layer

Dark layer

White layer

Dark layer

White layer

Dark layer

Fig. 9. White layers generated at: (a) 213.36 m/min (700 SFPM), 0.127 mm/rev (0.005 ipr), (b) 274.32 m/min (900 SFPM), 0.178 mm/rev (0.007 ipr), (c) 213.36 m/min (700 SFPM), 0.127 mm/rev (0.005 ipr), (d) 274.32 m/min (900 SFPM), 0.178 mm/rev (0.007 ipr).

2.5

White Layer Thickness, m

1.5

0.5

0 213/0.127 274/0.127 213/0.178 Speed (m/min) / Feed (mm/rev) Predicted (m) Measured Avg. (m) 274/0.178

Fig. 10. Comparison between experimental and predicted values of white layer thickness.

Author's personal copy


ARTICLE IN PRESS
410 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414

Eqs. (4)(11) describe the procedure for the simulation of phase transformation during machining of the steel. It must be noted that only diffusionless transformations, namely, martensite formation, are accounted for by the equations outlined above. The conditions used for the simulation of white layer formation are shown in Table 5. The FE simulation procedure is identical to that described earlier. A mesh using 3.125 mm long elements is used in order to maximize the sensitivity to temperature changes along the workpiece surface. The tool material is a low-cBN-content grade whose properties are given in Table 6. Note that the tool was modeled with a ank wear land width of 100 mm to match the conditions used in the model validation experiments described in the following paragraphs. The austenite and martensite contents were monitored using solution-dependent state variables dened in ABAQUSs. A sample plot showing the occurrence of phase transformation (indicated in red) for the 274.32 m/min and 0.178 mm/rev (900 SFPM/0.007 ipr) case is shown in Fig. 8. The unshaded elements immediately in front of the tool are elements in the sacricial layer used to model the chip separation region. Dry orthogonal cutting tests (see Fig. 3) were carried out on a 0.9 mm thick tube of hardened 52100 steel (62 HRC) on a two-axis CNC lathe (Hardinge Model T42SP) at these conditions using low-cBN-content tools (Kennametal NG3125L KD081) with articially generated ank wear lands of approximately 100 mm. The lower thermal conductivity of this tool (compared to high cBN) helps in channeling more heat into the chip-work region, thereby promoting white layer formation. At the high cutting speeds used in these tests, even a fresh tool rapidly develops a measurable ank wear, which contributes to frictional heating of the workpiece surface and thereby white layer formation. A 01 rake angle tool holder (Kennametal NER 123B NF1) was used in the tests. The workpiece was sectioned and samples from two locations along the machined length were polished and etched using standard metallographic techniques. For the purpose of measuring the white layer thickness, optical micrographs were generated at two locations per sample at a magnication of 2000. The white layer thickness was measured at 4050 locations per micrograph and averaged. The measurement error was estimated to be about 10%. The white layers generated at the different machining conditions used in the tests are shown in Fig. 9. Note that the somewhat darker region immediately below the white layer in the micrographs is the so-called dark layer, which mainly consists of tempered martensite. A comparison of experimentally determined white layer thickness and predicted values is shown in Fig. 10. Note that the predictions were made via model simulations that accounted for the 100 mm articially ground ank wear used in the experiments. It can be seen from the gure that better agreement between predicted and measured values of

white layer thickness is obtained at higher feeds. However, it is found that all predicted values lie within 1.5 times the standard deviation of the measured thickness. One potential explanation for the difference between the measured and predicted values of white layer thickness is the error associated with the measurement technique. Another explanation is the mesh density near the workpiece surface. Simulations were run with a mesh height of approximately 0.75 mm near the workpiece surface. Further renement would possibly improve some of the predictions, but the computational cost would increase disproportionately.

Fig. 11. Validation of residual stresses predicted by nite element simulations of orthogonal cutting. Experiment conditions reported by Guo [46]: 0.1016 mm feed, 0.762 mm depth of cut, 51 rake angle, 1.6 mm tool nose radius.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 411

Cutting tool inserts typically have a nominal cutting edge radius (or hone). Honed edges promote temperature rise at the toolwork interface. This effect was not considered in the FE simulations. Dynamic remeshing and rezoning techniques, which should alleviate problems associated with mesh size and cutting edge geometry, should improve the accuracy of the simulations. These features are currently unavailable in ABAQUSs v6.2. 5. Modeling of residual stresses In order to track residual stress evolution in the workpiece after machining, the following approach is used:

worked well for all the machining conditions studied. A sink temperature of 298 K (25 1C) is specied as the environment temperature. Model temperatures were found to be in the vicinity of 300 K (27 1C) after a simulation period of 1000 s. In order to verify the accuracy of the solution procedure, results from FE simulations are compared against experimental data for nose turning of AISI 52100 steel (62 HRC) reported by Guo [46]. The cutting conditions used in the three-dimensional nose turning test reported in Guos work are as follows: 106.68 m/min, 0.762 mm depth of cut and 0.1016 mm/rev feed. The tool (BZN 8100) used in this test had a nominal rake angle of 51 and a clearance angle of 51. Since much of the cutting in this test occurs around the tool nose, the undeformed chip thickness is non-uniform. Therefore, in order to achieve a meaningful comparison between the residual stress distribution predicted by the white layer model presented in this paper and the data reported by Guo, an effective undeformed thickness value that accounts for the effect of tool nose radius needs to be determined and used in the orthogonal FE simulations of the residual stress. The effective undeformed chip thickness in nose turning was calculated using the

 

Once the machining pass is completed, the workpiece is decelerated to a full stop in approximately 5 ms. The tool is then backed away from the workpiece to ensure that no surfaces of the tool are in contact with the workpiece. The tool is also brought to a complete halt. All thermo-mechanical loads applied to the model are then deleted and the workpiece is allowed to cool to room temperature. This is achieved by specifying a convection lm coefcient along the workpiece boundary. It is found that a value of 1.5 104 J/s mm2 K

Fig. 12. Effect of white layer formation on circumferential residual stress.

Author's personal copy


ARTICLE IN PRESS
412 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414

following equation [47]: teff rn cos1 rn dt , d=rn sin1 t=2rn (12)

where d is the depth of cut, t is the nominal undeformed chip thickness and rn is the tool nose radius. Using d 0.762 mm, t 0.1016 mm and rn 1.6 mm yields a value of 0.046 mm for the effective undeformed chip thickness. This value was used in the FE simulation of residual stress in orthogonal cutting of AISI 52100 steel. The width of cut in the simulation was equal to 0.762 mm. As seen in Fig. 11, a fairly good agreement is obtained between the predicted and measured values of the circumferential and axial residual stresses. 5.1. Effect of white layer formation on residual stresses In order to illustrate the effect of white layer formation on the residual stresses, machining simulations for the conditions listed in Table 5 were run with and without the white layer model described previously. The simulated results are shown in Figs. 12 and 13. It can be seen that the model simulations accounting for white layer formation yield surface residual stresses that are more compressive than when white layer formation is

not included. This is attributed mainly to the strains caused by volume expansion of the material during phase transformation. This would actually serve to offset the tensile residual stresses caused by thermal input to the workpiece. It can also be seen from Figs. 12 and 13 that the location of the peak compressive stress appears to shift toward the workpiece surface when white layer formation is considered in the simulations. Specically, it is found that the peak compressive stresses occur at depths of approximately 1020 mm below the workpiece surface. This trend is consistent with experimentally determined values for hardened steel workpieces [48,49]. These results suggest that accounting for white layer formation in the simulation of hard machining will result in calculated residual stress proles that are closer to experimentally determined values in a qualitative sense. 6. Conclusions The paper presented an FE model of continuous white layer formation in orthogonal cutting of AISI 52100 hardened steel. The problem is likened to a quenching process and is modeled as such. The model explicitly incorporates the effects of stress and strain on the

Fig. 13. Effect of white layer formation on axial residual stress.

Author's personal copy


ARTICLE IN PRESS
A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 413

transformation temperature, volume expansion and transformation plasticity and is valid for continuous white layer formation under thermally dominant cutting conditions where phase transformation is likely to occur. Experimental validation of the model is shown to yield predicted values and trends of white layer thickness that are in good agreement with the measured values and trends. It is found that accounting for white layer formation in the model simulations results in more compressive surface residual stress, which is explained by the volume expansion accompanying the phase transformation. It is also found that accounting for white layer formation causes the location of the peak compressive stress to shift toward the workpiece surface, as has been observed experimentally by others [48,49]. Acknowledgments This work was supported by a grant from the National Science Foundation (DMI-0100176). Additional support was provided by the NIST ATP Program (#70NANBOH3045). The authors are also grateful to the National Center for Supercomputing Applications (NCSA) at the University of Illinois at Urbana-Champaign for providing computer time for running the ABAQUS simulations. References
[1] S. Akcan, S. Shah, S.P. Moylan, P.N. Chhabra, S. Chandrasekhar, H.T.Y. Yang, Formation of white layers in steels by machining and their characteristics, Metallurgical and Material Transactions A 33A (2002) 12451254. [2] A. Barbacki, M. Kawalec, Structural alterations in the surface layer during hard machining, Journal of Materials Processing Technology 64 (1997) 3339. [3] A. Barbacki, M. Kawalec, A. Hamrol, Turning and grinding as a source of microstructural changes in the surface layer of hardened steel, Journal of Materials Processing Technology 133 (2003) 2125. [4] M. Field, J.F. Kahles, Review of surface integrity of machined components, Annals of the CIRP 20 (2) (1971) 153163. [5] D.M. Turley, The nature of the white-etching layers produced during reaming ultra-high strength steel, Materials Science and Engineering 19 (1975) 7986. [6] H. Eda, K. Kishi, S. Hashimoto, The formation mechanism of ground white layers, Bulletin of the JSME 24 (190) (1981) 743747. [7] B.J. Grifths, Mechanisms of white layer generation with reference to machining and deformation processes, Transactions of the ASME: Journal of Tribology 109 (1987) 525530. [8] H.K. Tonshoff, H.G. Wobker, D. Brandt, Hard turninginuences on the workpiece properties, Transactions of NAMRI/SME XXIII (1995) 215220. [9] A.M. Abrao, D.K. Aspinwall, The surface integrity of turned and ground bearing steel, Wear 196 (1996) 215220. [10] E. Brinksmeier, T. Brockhoff, White layers in machining steels, in: Proceedings of the Second International German and French Conference on High Speed Machining, 1999, pp. 713. [11] A. Vyas, M.C. Shaw, The signicance of the white layer in a hard turned steel chip, Machining Science and Technology 4 (1) (2000) 169175. [12] J.D. Thiele, S.N. Melkote, T.R. Watkins, R.A. Peascoe, Effect of cutting edge geometry and workpiece hardness on surface residual stresses in nish hard turning of AISI 52100 steel, Transactions of the

[13]

[14]

[15]

[16]

[17]

[18]

[19]

[20]

[21] [22]

[23]

[24]

[25]

[26]

[27] [28] [29]

[30]

[31]

[32]

ASME, Journal of Manufacturing Science and Engineering 122 (4) (2000) 642649. J. Barry, J. Byrne, TEM study on the surface white layer in two turned hardened steels, Materials Science and Engineering A 325 (2002) 356364. X. Sauvage, J.M. Le Breton, A. Guillet, A. Meyer, J. Teillet, Phase transformations in surface layers of machined steels investigated by X-ray diffraction and Mossbauer spectrometry, Materials Science and Engineering A 362 (2003) 181186. V. Mahajan, Y.K. Chou, Machining-induced surface hardening due to large wear land, Proceedings of the Institution of Mechanical Engineers, Part B: Journal of Engineering Manufacture 218 (2004) 16471655. Y. Guo, J. Sahni, A comparative study of hard turned and cylindrically ground white layers, International Journal of Machine Tools and Manufacture 44 (2004) 135145. A. Ramesh, S.N. Melkote, L.F. Allard, L. Riester, T.R. Watkins, Analysis of white layers formed in hard turning of AISI 52100 steel, Materials Science and Engineering A 390 (2004) 8897. Y. Huang, S.Y. Liang, Modeling of cutting forces under hard turning conditions considering tool wear effect, Transactions of the ASME, Journal of Manufacturing Science and Engineering 127 (2) (2005) 262270. Y.B. Guo, C.R. Liu, 3D FEA modeling of hard turning, Transactions of the ASME, Journal of Manufacturing Science and Engineering 124 (2) (2002) 189199. T.D. Marusich, R.J. McDaniel, S. Usui, J.A. Fleischmann, T.R. Kurfess, S.Y. Liang, Three-dimensional nite element modeling of hard turning processes, Proceedings of the ASME Manufacturing Engineering DivisionMED 14 (2003) 221227. Y. Huang, T.G. Dawson, Tool crater wear depth modeling in cBN hard turning, Wear 258 (9) (2005) 14551461. M. Mahdi, L. Zhang, Correlation between grinding conditions and phase transformation of an alloy steel, in: L.C. Wrobel, et al. (Eds.), Advanced Computational Methods in Heat Transfer III, Computational Mechanics Publications, Southampton, 1994, pp. 193200. M. Mahdi, L. Zhang, Applied mechanics in grindingVI. Residual stresses and surface hardening by coupled thermo-plasticity and phase transformation, International Journal of Machine Tools and Manufacture 38 (1998) 12891304. M. Mahdi, L. Zhang, Applied mechanics in grinding. Part 7: residual stresses induced by the full coupling of mechanical deformation, thermal deformation and phase transformation, International Journal of Machine Tools and Manufacture 39 (1999) 12851298. N.S. Akcan, Microstructure analysis and temperature modeling in nish machining of hardened steels, M.S. Thesis, School of Industrial Engineering, Purdue University, 1998. Y.K. Chou, C.J. Evans, White layers and thermal modeling of hard turned surfaces, International Journal of Machine Tools and Manufacture 39 (1999) 18631881. D.A. Porter, K.E. Easterling, Phase Transformations in Metals and Alloys, Van Nostrand Reinhold, New York, 1981. G. Krauss, Steels: Heat Treatment and Processing Principles, ASM International, Metals Park, OH, 1990. K.F. Wang, S. Chandrashekhar, H.T.Y. Yang, Experimental and computational study of the quenching of carbon steel, Transactions of the ASME, Journal of Manufacturing Science and Engineering 119 (1997) 257265. F.G. Rammerstofer, D.F. Fischer, W. Mitter, K.J. Bathe, M.D. Snyder, On thermo-elasticplastic analysis of heat treatment processes including creep and phase changes, Computers and Structures 13 (1981) 771779. S. Denis, E. Gautier, A. Simon, G. Beck, Stress-phase transformation interactionsbasic principles, modelling, and calculation of internal stresses, Materials Science and Technology 1 (1985) 805814. S. Denis, S. Sjostrom, A. Simon, Coupled temperature, stress, phase transformation calculation model numerical illustration of the

Author's personal copy


ARTICLE IN PRESS
414 A. Ramesh, S.N. Melkote / International Journal of Machine Tools & Manufacture 48 (2008) 402414 internal stresses evolution during cooling of a eutectoid carbon steel cylinder, Metallurgical Transactions A 18A (1987) 12031212. S. Denis, E. Gautier, S. Sjostrom, A. Simon, Inuence of stresses on the kinetics of pearlitic transformation during continuous cooling, Acta Metallurgica 35 (7) (1987) 16211632. K.F. Wang, S. Chandrashekhar, H.T.Y. Yang, An efcient 2-D nite element procedure for the quenching analysis with phase change, Transactions of the ASME, Journal of Engineering for Industry 115 (1993) 124138. A.K. Jena, M.C. Chaturvedi, Phase Transformations in Materials, Prentice-Hall, Englewood Cliffs, NJ, 1992. Z. Nishiyama, Martensitic Transformations, Academic Press, New York, 1978. G.R. Johnson, W.H. Cook, in: Proceedings of the Seventh International Symposium on Ballistics, American Defence Preparedness Association (ADPA), Netherlands, 1983. M. Shatla, C. Kerk, T. Altan, Process modeling in machining, part I: determination of ow stress data, International Journal of Machine Tools and Manufacture 41 (2001) 15111534. M. Baker, C. Siemers, J. Rosler, A nite element model of high speed metal cutting with adiabatic shearing, Computers and Structures 80 (56) (2002) 495513. Y.B. Guo, D.A. Dornfeld, Finite element modeling of burr formation process in drilling 304 stainless steel, Journal of Manufacturing Science and Engineering 122 (2000) 612619. [41] A. Ramesh, Prediction of process-induced residual stresses and microstructural changes in orthogonal hard machining, Ph.D. Thesis, Mechanical Engineering, Georgia Institute of Technology, 2002. [42] R.L. Martineau, A viscoplastic model of expanding cylindrical shells subjected to internal explosive detonations, Ph.D. Thesis, Department of Mechanical Engineering, Colorado State University, Fort Collins, CO, 1998. [43] L.S. Darken, R.W. Gurry, Physical Chemistry of Metals, CBS Press, New Delhi, 1953. [44] J. Beswick, Effect of prior cold work on the martensite transformation in SAE 52100, Metallurgical Transactions 15A (1984) 299306. [45] D.P. Koistinen, R.E. Marburger, A general equation prescribing the extent of the austenitemartensite transformation in pure iron carbon alloys and carbon steels, Acta Metallurgica 7 (1959) 59. [46] Y.B. Guo, 3-D modeling of supernish hard turning, Ph.D. Thesis, Department of Industrial Engineering, Purdue University, 2001. [47] T.G. Dawson, Effects of cutting parameters and tool wear in hard turning, Ph.D. Thesis, School of Mechanical Engineering, Georgia Institute of Technology, 2002. [48] E. Schreiber, H. Schlicht, Residual stresses after turning of hardened components, in: G. Beck, S. Denis, A. Simon (Eds.), Proceedings of the Second International Conference on Residual Stresses (ICRS 2), Elsevier, New York, 1988, pp. 853860. [49] A. Ramesh, J.D. Thiele, S.N. Melkote, Residual stresses and subsurface ow in nish hard turned AISI 4340 and 52100 steels: a comparative study, Proceedings of the ASME IMECE, MED 10 (1999) 831837.

[33]

[34]

[35] [36] [37]

[38]

[39]

[40]

Vous aimerez peut-être aussi