"‘fihm‘w‘ I fixaéiii‘ékfi' <‘.- Iv up 31;, ,3 .. m: « v0 .7 w‘ '9‘ .r ‘ q... I -, "W ’. 1 - Ara. ’ ' I! AN STA minimumIMHII’H‘IIWI’I‘HWWIWI’II‘WI 31293 017141502 This is to certify that the thesis entitled PREDICTIVE TOOL WEAR OF COATED TOOL SYSTEMS presented by Raja Krishnan Kountanya has been accepted towards fulfillment of the requirements for M. S . . Mechanics degree in MW Major professor Date g/ci-l/yc?“ 076% MS U i: an Affirmative Action/Equal Opportunity Institution LIBRARY Mlchlgan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINE return on or before date due. MTE DUE MTE DUE J. 1/98 chlRClDahDuapGS—p.“ PREDICTIVE TOOL WEAR OF COATED TOOL SYSTEMS By Raja Krishnan Kountanya A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Materials Sciences and Mechanics 1998 ABSTRACT PREDICTIVE TOOL WEAR OF COATED TOOL SYSTEMS By Raja K.Kountanya Tool wear has been a serious concern in the economics of modern machining. In the light of the developments in materials technology, the demand for suitable cutting tool materials has entailed a detailed study into the various mechanisms that bring about tool wear. Temperature has been known to play an important role in the mechanisms of wear of cutting tools. The literature is rich in studies that focus purely either on the temperature fields or the mechanisms that bring about tool wear. This study adopts a dual approach to the problem by empirical quantification of cutting temperatures and semi-analytical modeling of tool wear so as to eliminate the need for laborious testing for optimal tool materials. Abrasion [Rabinowicz, 1961] and chemical dissolution [Kramer and Suh, 1980] are understood to dominate the wear process of the tool for the work materials studied. Experimentally obtained wear data and that predicted theoretically, brought to light many interesting aspects in the tool wear problem. Cutting tests were conducted on plain carbon steels of A181 designation 1018, 1045, 1065, 1070 and 1095 with carbide cutting tools with a single coating of TiN, TiCN and A1203. Temperature of the cutting tool was measured using an infrared pyrometer with a fiber optic attachment. An inverse estimation was then carried out to estimate the interface temperatures. Flank wear rate increased with cementite content. Crater wear rate increased with temperature attesting to the common notion that thermally activated wear mechanisms brings about crater wear. In general, good correlation between experiment and theory was found. ACKNOWLEDGEMENTS I wish to place on record my sincere thanks to my advisor Dr.Patrick Youngseon Kwon, for rendering all the technical and moral support during the course of this work. I am deeply indebted to his enduring patience and understanding, which was invaluable towards bringing this study to fruition. My thanks to Thorsten Schiemann, from Aachen, Germany, whose insightful suggestions and practical knowledge clarified many a doubt, which arose during the experiments and the data processing. I consider my association with him as a delightful experience, which taught me the essence of teamwork. I am indebted to Ben Simkin, a fellow graduate student for his valuable time in obtaining the images on the Scanning electron microscope. My thanks to my thesis committee members, Dr.Thomas Bieler and Dr.K.N.Subramaniam for their thoughtful suggestions in the writing of this thesis. My thanks to Mr.Fritz Smydra of Lansing Community College for his expertise on the CNC lathe. His availability on some crucial weekends is deeply appreciated. My thanks also to Kurt Niemeyer and Mr.J.C.Brenton of the machine shop in the College of Engineering, M.S.U. for 'lending me a hand ' on many an occasion. Last but not the least, my thanks to all my friends in and around campus, including Biman Ghosh, Sunil Chandran, Suhail Ansari, Monica Lokre and Sridhar Seshagiri whose encouragement was the 'magic tonic' to keep me going. The moral I gained from the whole experience was the simple truth in the sanskrit saying ‘Siddhirbhavati Karmaja’, meaning ‘Do Unto Perfection’. iii TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES LIST OF SYMBOLS OVERVIEW CHAPTER 1 INTRODUCTION TO TOOL WEAR 1.1 Geometrical aspects of tool wear 1.2 Literature Survey in relation to tool wear 1.3 Importance of Temperature in relation to too] wear CHAPTER 2 STATEMENT OF THE PROBLEM 2.1 Rabinowicz Three-body and Two-body wear models 2.2 Chemical Dissolution wear of Kramer and Sub 2.3 Comprehensive models for Crater and Flank Wear 2.4 Accounting the variation of hardness with temperature 2.5 Comparison with Taylor’s model for tool wear iv vii viii \O-bNN 1o 10 12 14 15 16 CHAPTER 3 CUTTING TOOL TEMPERATURES 3.1 Inverse Problem of measuring the interface temperatures 3.2 1D Ellipsoidal model of Yen and Wright 3.3 Oxley’s method of obtaining the flank temperature CHAPTER 4 EXPERIMENTAL WORK 4.1 Turning Experiments 4.1.1 Cutting Conditions and related issues 4.2 Method of temperature measurement 4.2.1 Principles of Infrared Pyrometry 4.2.2 Implementation of the Infrared Pyrometer for the Experiments 4.3 Data Acquisition 4.4 Work Materials 4.4.] Dimensions of the bar-stock, cutting length and duration of cuts 4.4.2 Spherodize-annealing of the steels 4.4.3 Microstructures, Composition and Hardness of the steels 4.5 Inserts used in the experiments 4.5.1 Insert Geometry and ISO/AN SI designation 4.5.2 Grade of substrate used 4.5.3 Coating Materials and related details 4.5.4 Designation for the identification of the inserts 4.6 Area Measurement 4.7 Tool Wear Measurement : set-up for Flank Wear and Crater Wear CHAPTER 5 RESULTS AND DISCUSSIONS 5.1 Results of the Computer Algorithm 5.1.1 Abrasive Wear 17 18 18 21 22 22 23 23 24 24 27 28 28 28 29 31 31 33 34 34 38 38 40 4o 42 5.1.2 Dissolution Wear 5.2 Temperature trends with cutting speed 5.3 Contact Area Trends 5.4 Flank Wear 5.5 Crater Wear 5.6 Calibration CHAPTER 6 SUMMARY AND CONCLUSIONS APPENDICES BIBLIOGRAPHY vi 42 44 45 47 52 54 57 59 91 Table 4.1: Table 4.2: Table 4.3: Table 4.4: Table 4.5: Table 4.6: Table 4.7: Table 5.1: Table 5.2: Curve fitted relations for the dissolution wear of the coating materials (x is temperature in °K, w.r. stands for wear per unit time) Table A]: Table A2: Table A3: Table A4: Table A5: Table A6: LIST OF TABLES Composition of the hot-rolled steels used in the testing (All in wt%) Composition of the spherodized steels used in the testing (All in wt%) Cutting tool signature Properties of the substrate Therrnophysical property data of substrates Process parameters for the spherodized steels testing Process parameters for the unspherodized steels testing List of wear rate equations for the calibration purpose (T is in °C, w.r. stands for wear per sliding distance) mechwear.out for the TiN coating mechwear.out for the TiCN coating mechwear.out for the A1203 coating chemdiss.out for the TiN coating chemdiss.out for the TiCN coating chemdiss.out for the A1203 coating vii 29 29 32 33 33 36 37 42 73 73 73 74 74 74 Figure 1.1: Figure 2.1: Figure 2.2: Figure 2.3: Figure 3.1: Figure 3.2: Figure 4.1: Figure 4.2: Figure 4.3: Figure 4.4: Figure 4.5: Figure 4.6: Figure 4.7: Figure 4.8: Figure 4.9: LIST OF FIGURES Zones of damage of the cutting tool due to machining Illustration of two-body and three-body abrasion Oversimplified model of a conical abrasive wearing a bearing surface to illustrate abrasive action Schematic illustrating chip flow and continuity conditions [Kramer and Suh, 1980] Sources of heat generation in a cutting tool Temperature distribution in a cutting tool for the inverse temperature estimation Photograph of the lathe with the experimental set-up (Inset: pyrometer end probe) Illustration of the chip-breaker set-up over the insert Final assembly of the pyrometer over the insert Microstructures of the steels used in the testing Hardness of the steels used in the machining tests Illustration of the various geometric details of the cutting tool Sample-naming of the TiN insert used on A181 1045 (spherodized) for the low speed cut. Illustration of the method for the area calibration Set-up on the TMM for measuring flank and crater wear Figure 4.10: Sample of the crater profiles (Cut-ID xoa1018sp-2) viii 10 12 13 17 18 22 26 27 30 31 32 35 38 39 39 Figure 5.1: Charts illustrating the ranking of the coatings for the two abrasive wear mechanisms Figure 5.2: Figure illustrating the ranking of the coatings for the dissolution wear mechanism Figure 5.3: Temperature records of cuts on AISI 1095 steel with TiN coated K420 carbide (Curve fitting shown alongside) Figure 5.4: Photographs of K420 carbide coated with TiN, TiCN and A1203 respectively afier machining spherodized AISI 1045 steel at increasing cutting speeds at f = 0.356 mm/rev and doc. = 1.905 mm. Figure 5.5: SEM micrographs (800x) of the worn flank surfaces of the coated tools after machining spherodized AISI 1045 steel for low and high speeds. Figure 5.6: Plots showing the variation of the volume flank wear per sliding distance for the three coating materials. (Sp-spherodized steel, un-unspherodized steel) Figure 5.7: Plots showing the variation of the crater wear rate for the three coating materials. (Sp-spherodized steel, un-unspherodized steel) Figure 5.8: Plots of the predicted and the experimental wear data for A1203 while machining spherodized steels. Figure 5.9: Plots of the constants obtained from the calibration process. (sp — spherodized steel, un — unspherodized steel) ix 41 43 44 46 49 51 53 55 56 KT VB VBmax V,Vm P1 is manna LIST OF SYMBOLS Crater width (um) Crater depth (um) Crater center distance (um) Crater area (mmz) Average wear land width (um) Maximum wear land width (um) Nose radius of insert (mm) Volume of worn material Roughness angle of the abrasive (deg) Sliding distance (m) Normal force of interaction between the surfaces (N) Tool material hardness (Kg/mm2) Abrasive Hardness (Kg/mmz) Calibration constant for abrasive wear per sliding distance Calibration constant for dissolution wear Solubility (dimensionless) Molar volume (cm3/mol) Gibbs’ free energy (Kcal/mol) >6 x,y, 2 B: Excess free energy of solution (Kcal/mol) Universal gas constant (kcal/mol.°K) Stoichiometric coefficients Volume worn away on the relief face (mm’) Flank wear land (mm) Constant for the softening (/°C) Hardness (Kg/mm2) Temperature (°C) Dimensionless temperature Distance in elliptical coordinates (mm’) Parameters describing the base ellipse (mm) Complete elliptical integral of the first kind Steady-State cutting interface temperature (°C) Temperature in the far-field of the tool (Taken to be 25°C) Molar Volume (cm3/mol) xi OVERVIEW Metal cutting has been a field of study of immense commercial importance. Since even minor improvements in productivity can result in enormous cost savings and profits in high volume production, there has been a never-ending demand for research in this field even after ten decades of industrialization. In the light of the enormous amount of work done in the past and the work going on presently, the scope of research has become very focussed. Hence, study of any problem in metal removal processes involves careful design of experiments, instrumentation and evaluation of results. A solid infrastructure is therefore necessary for in this field. Development of cutting tool materials has been foremost in all the research in related to metal cutting. A significant stride in this regard has been the innovation of adopting certain coating technologies for depositing thin hard coatings on common tool substrates. This technique, developed in the early 19705, dramatically improved tool life and productivity. Today more than 75% of turning operations and 40% of milling and drilling operations are performed on coated carbides. While hot-hardness and chemical inertness have been recognized to be the two important parameters for coating materials, the exact dependencies and the rationalization are yet to take a concrete shape. Chapter 1 INTRODUCTION TO TOOL WEAR Cutting edge wear is still one of the unsolved problems in metal cutting. No matter how superior the tool material is, it is common experience that tool life, defined as the time for which tool wear is within acceptable limits, is always finite. Consequently, the reasons for interest in tool wear are threefold: (1) Lower workpiece quality due to deterioration in finish and dimensional accuracy, or damage to expensive workpieces if an edge fails catastrophically in a cut. The second is (2) the cost of changing cutting edges, the cost associated with the concomitant time delays and damage caused by unexpected failure and (3) the increased power consumption due to the excessive rubbing of the tool with the work material. Work continues for three basic purposes: (1) how to accurately predict wear, (2) how to detect wear from measurements while machining, and (3) how to minimize wear by the development of new materials. This study was focussed on the first purpose. 1.1 Geometrical Aspects of Tool Wear In order to characterize tool wear the proper geometric parameters have to be defined. The ISO standard 3685-1977(E) [1] for tool life testing was formulated with this express purpose so that there can be a basis for comparison between various cutting tool materials. Figure 1.1 illustrates the geometric parameters associated with tool wear in a typical right hand turning tool. Section A—A Tool face V crater area KA KB Kl flear notch KB= crater width Kl=crater centre distance KT=crater depth KA=crater area (self defined) VB=average wear-land width VBnax=maximum wear-land width r=radius of cutting edge View on major flank Figure 1.1: Zones of damage of the cutting tool due to machining 1.2 Literature Survey in relation to tool wear In the light of the numerous attempts to investigate tool wear, the following section will detail briefly the important information gathered from the recent literature concerning this study. Ramalingarn and Wright [2] performed studies on well-characterized work materials such as F e-C-silica powder-metallurgy compacts where they found flank wear to increase with silica content. They concluded that alloy chemistry did not describe machinability and found the wear process on the tool flank to be dependent on interfacial temperatures. The prowing action of the abrasive particles was clearly shown by means of quick-stop sections. They also concluded that while the tool material could soften considerably at the temperature prevalent at the interfaces, the abrasive particles did not soften by the same amount since the duration for which they were present in the shear zone was extremely small, of the order of milliseconds. Tool chip temperatures were higher for carbide tools due to the higher speeds used on them. A modest assumption that the flank temperature was 300-400°C lower than the chip-tool interface temperature was made. Byrd and Ferguson [3] studied the influence of hard inclusions on flank wear of carbide tools. A1203 particles were artificially impregnated in 1020 and 4620 steels using the PM technique. They concluded that higher temperatures encountered in the machining of steels do not in any way impede the abrasive action of hard inclusions in the microstructure. However, they also had difficulties in establishing a threshold level in the degradation of machinability due to hard particles in the microstructure. Vol.% of hard particles seemed to provide a clue to the machinability. Brun et a1. [4] performed experiments on 40%vol. SiC aluminum alloy with different tool materials. Since they were cutting aluminum, cutting temperatures encountered were expected to be low. They found that tool materials harder than SiC performed much better than the others. This dependence on hardness represented a complex behavior wherein abrasion was dependent on temperature at the interfaces and the altered properties of the cutting tool at this temperature. Kramer [5] investigated the wear resistance of binary carbide coatings. The WC substrates began deforming thermally nearly at 1030K and showed a rapid decline in the compressive strength. At higher cutting temperatures, he concluded that chemical dissolution was the most important wear mechanism. Ramalingam and Watson [6] considered the factors responsible for the scatter in tool life. The role and significance of the oxygen rich non-metallics on tool wear and machinability was examined. Tool chip interface temperatures of carbide tools were 800°C and above under normal industrial practices. Diffusion was excluded as a contributing wear mechanism since it was thought that diffusion was a well behaved, non-equilibrium mechanism that would lead to a deterministic solution for tool life. In carbide tooling the ‘prowing’ process was believed to give rise to plastic flow in the carbide at the tool-chip interface. Hence, a plausible reason for the stochasticity was the variation in distribution of the abrasive particles in the work material. Kramer and Kwon [7] concluded that tool wear was primarily due to two mechanisms namely chemical dissolution and abrasion. The dissolution of the tool material in the moving stream of chip material may be treated as a dilute solid solution formation and can be modeled as a regular solution. They also concluded that even though wear models such as the 3-body and 2-body abrasion do not describe the constraints on the abrasive particle in machining, they can be adopted for modeling abrasive wear. The ferrite matrix was believed to be quite soft at the temperatures present in the zone of deformation during cutting, offering minimal restraint to the inclusions. The details of two-body and three-body abrasion, an important highlight of this study, will be presented later on in this thesis. Kramer and Sub [8] developed the dissolution wear model, which will be the topic of discussion later on. They concluded that solution wear is predominant in the carbide class of tool materials, which is independent of the diffusion characteristics. In all cases the tool material was assumed to be dissolving in a-iron. The solubility in y-iron was excluded due to the commonly observed sluggishness of the 01—97 transition. Essentially, this study proved that the free energy of formation of the ceramic coating determined the effectiveness of the coating. The hypothesis was confirmed from the ranking obtained from the experimental results of crater wear and that obtained from the thermodynamic calculations. The difference of this model from the formerly believed diffusion theory is that dissolution is an equilibrium process whereas diffusion is not. Dearnley [9] cut'various plain carbon steels carbide tools with a single or composite layers of TiC, TiN and A1203. Cutting temperature was measured using the tool-work thermocouple method and was further verified using the metallurgical technique. Uncoated tools showed a larger HAZ (Heat Affected Zone). The temperature difference between the uncoated tools and the coated tools was found to be less than 150°C. Although clear evidence of seizure was noted on the rake surface, there were indications of reduced interfacial contact on the flank. The A1203 coatings and the ceramic inserts showed the greatest propensity for ridges on the rake face via discrete plastic deformation giving further evidence to abrasion on the flank since temperatures are lower on the flank. This was the basis for the conclusion that wear of A1203 was primarily due to a decohesion mechanism. Moreover, flank wear trends never followed the trend indicated by the dissolution or diffusion mechanisms indicating that they may not be the rate controlling mechanisms, i.e. the mechanisms contributing the most to tool material removal, on the flank. However, they were useful in interpreting the preferential dissolution of the WC substrate. Cho and Komvopoulos [10] studied wear mechanisms of multi-layered coated tools. Severe abrasion was noted at the flank because of the lower temperature than the crater, the more rigid work material and the constraint of the moving work and the tool. In the case of A1203, they noted that dissolution may be neglected at all cutting speeds and mechanisms such as plastic flow, thermo-mechanical fatigue and fracture were expected to prevail at all temperatures. Cook [11] observed that the average wear land temperature seemed to approach the tool-chip interface temperature as tool wear progresses. At higher speeds, crater wear rates were primarily a function of the temperature. Functionality of flank wear with temperature was not deducible precisely. Stj emberg and Thelin [12] noted that increasing the coating thickness increased the overall resistance to crater wear. The time needed to expose the substrate underneath was the determining factor in this regard. It was suggested that the temperatures in the flank are 300°C lower than that at the chip-tool interface. All coatings were harder and more ductile at lower temperatures. Notch wear was noted to be a chemical phenomenon. Kramer [13] suggested that at moderate cutting temperatures, excessive rubbing occurs between the flank and the work material. F lank wear determined tool life at low speeds due to mechanically activated wear caused by microfracture, thermal and mechanical fatigue and abrasion by hard inclusions. Milovic et a1. [14] noted that the fluctuating stress conditions that can exist within the BUE in the machining of free cutting steels can be the reason for the superior performance of HSS tools as against carbide tools, owing to their superior toughness. The coating reduced the interface temperature by as much as 125°C and hence could be used to turn the material at a speed higher by 25 m/min. The thermal conductivity of carbide tools, which is higher than that for HSS, was another reason for the higher heat abstraction from the interfaces and hence higher speeds needed for machining without a BUE. Chubb and Billingham [15] studied the wear mechanisms in high speed machining. They found that once the coating was removed from the flank, the mechanism of wear appeared to be a combination of abrasion and diffusion. WC appeared to wear by diffusion as evidenced by the smooth boundary between the WC particles and the steel. Hence tool life due to flank wear was closely related to the breakdown of the coating on the flank. Kim [16] noted that abrasive wear dominated the wear on the flank surface and diffusion on the crater surface. He also concluded that excessive coating thickness could retard wear performance. Kim and Durham [17] noted that cutting temperature as high as 1600°C could be reached with alumina tools without tool failure. The temperature at the flank could be 100°C lower. Tools with a higher thermal conductivity and a higher hot hardness showed a higher resistance to flank wear. Lee and Richman [18] noted that coated tools resist cratering even after the coating had been removed in some places. Hardness of a material as a coating was noted to be very different from that as a bulk material. Cooling of CVD coatings after deposition at 1000°C was noted to develop tensile residual stresses in them. Hardness measurements on coatings were particularly difficult because of the smooth surfaces that are demanded for accurate measurements. Suh [19], in his classic paper, outlined the essential ingredients for the making of high performance coatings. Both mechanically and chemically activated wear processes were pronounced to depend sensitively on temperature. In general, though tools should be at least 4-4 1/2 times harder than the work material, this is not applicable to coatings since crater wear rates and hardness of the coatings did not correlate very well. Residual stresses existed in the coatings and were a function of the CTE differences between the coating and the substrate. An important point made in connection with this study is that contrary to A1203, SiC or Si02, Fe3C has a very high free energy of formation and is likely to dissociate at high temperatures. There has however, been no experimental evidence to date in this regard. He also mentions that A1203 coatings may have problems adhering to the substrate since A1203 did not permit diffusion of carbon atoms across the interface which relates to the common experience of depositing an intermediate layer of an adherent material like TiN, as was the case in this study. Subrarnaniam et a1. [20] performed high speed machining on AISI 1045 steel. They noted that crater area increased with cutting speed. Evidence of twinned martensite in the chips quenched in water were in accord with the average interface temperature predictions for the secondary shear zone since steel undergoes a martensitic transformation at these temperatures. The thermodynamic potential for dissolution is the most important criterion for the design of a coating to minimize crater wear at high speeds. Trent [21,22,23] is one of the pioneers in recognizing the exact interfacial conditions on the rake face. He was also the first to perceive the importance of recognizing the mechanisms that control tool wear. In his 1963 series of papers, he was the first to depart from the then accepted notion of the interfacial conditions. It was then believed that there was relative motion between the chip and the tool on the rake face. It has now been established beyond doubt that chip flow resembles fluid flow with an initial region where there is relative movement and thereafter there is complete seizure thereby forming the 'secondary' shear zone. Among the other pioneers in the field are G.Boothroyd, P.L.B.0xley, 0.0ptiz and M.C.Shaw. Their innumerable contributions to the literature have enabled a very scientific understanding of metal cutting today. 1.3 Importance of Temperature in relation to tool wear From the above section the importance of cutting tool temperatures in tool wear can be appreciated. Not only does a higher temperature bring about softening of the tool but it also makes the cutting tool vulnerable to thermally-activated mechanisms such as dissolution and diffusion. It can also be seen from the literature survey that only a few attempts have been made at a thorough quantitative study of tool wear involving temperatures. The present study will create precedence for one in the future so as to enable a very scientific and rational framework for predictive tool wear. Chapter 2 STATEMENT OF THE PROBLEM The principal aim of this study is, as mentioned before, is a quantitative study of tool wear involving cutting tool temperatures. Among the various mechanisms, which have been presented in the literature, two main mechanisms namely, abrasion and chemical dissolution are selected and formulated for the purpose. A calibration scheme is then proposed wherein tool wear can be predicted for a given cutting temperature. 2.1 Rabinowicz Three-Body and Two-body Abrasive Wear Models Among the various abrasive wear models in tribology literature, the one appropriate to modeling abrasive tool wear is the three-body abrasive wear model. Rabinowicz et al. [24] performed experiments wherein two surfaces slid against each other with the abrasives introduced in-between [Fig.2.1]. They drew conclusions related to wear rates, sliding conditions and material hardness. A number of materials and abrasives were chosen and a general empirical relation was found to fit the data. The final form of the equations, as applicable to tool wear, is as shown in equations 2.1. z-aouyAimon J—Mylhmsim Q Figure 2.1: Illustration of two-body and three-body abrasion. ll thanB V” = , fl" < 0.8 31: I: ~25) = xL “116(5) , 1.25 > 5- > 0.8 Equations. 2.1 "' 5.3Pt If, I; = than19[fL] , 5‘ > 1.25 "’ 2.43Pt If, P a where tan 6? is the average tangent of roughness angle of the abrasive grains (3 measure of the particle shape or sharpness), x is a sliding distance, L is the normal force of interaction between the surfaces, 17 is the hardness of tool and If, is the hardness of the abrasive. Eqs. 2.1 calculate the abrasive wear volume as a function of a sliding distance, x. In theory, the equations always governed the volume of material removed, unless the process of abrasion was preceded by the formation of cracks. In the presence of cracks, material is removed by brittle fracture as well as by abrasive wear. Since the tool wear pattern usually seen on the flank, which consists of characteristic groove marks, resembles a plastic ploughing process, the empirical relations can be expected to describe three-body abrasion in tool wear also. While three-body conditions exist when two bodies slide against each other with the simultaneous rolling of a hard abrasive particle in-between, two-body conditions represent a hard surface sliding over a relatively soft surface. The relations for two-body wear are relatively simpler. The volume removed per sliding distance is expressed in equation 2.2. d_V_Ltan6 dl_ 72p Equation 2.2 I where L = Load between interacting surfaces 12 0 = angle for indentation of the conical abrasive particles p, = hardness of the abraded surface, here the tool flank. It should be noted that, in this model, the hardness of the abrasive does not appear in the relation for the wear per sliding distance. Also the hardness of the abrasive is assumed to be infinite. Figure 2.2 shows the schematic used to obtain the relation. V p ‘ fig-‘P—it’: V \\\ /// I / ~ oume \ \ \\\\\\\V //Boarin¢ 4 d! > Figure 2.2 Oversimplified model of a conical abrasive wearing a bearing surface to illustrate abrasive action 2.2 Chemical Dissolution Wear of Kramer and Sub Kramer and Sub [8] treated crater wear as an equilibrium process of dissolution and that a thermodynamic potential existed for the solution of the tool material in the chip material. They obtained excellent correlation between the wear rates predicted by theory and that from experiment. This is in opposition to the thermal process of diffusion, which was then widely believed to be the cause of crater wear, since diffusion is a non- equilibrium rate process. Diffusion kinetics [8] were proved to be relatively slow at normal cutting temperatures. Figure 2.3 shows the schematic used to describe the phenomenon. The solution wear rate, in terms of the solubility and other parameters, is given by equation 2.3. Chemical Dissolution Wear Rate = [BMCVO'S] Equation 2.3 13 where M is the molar volume, C is chemical solubility of the tool material in the work material, V is the cutting velocity and B is a calibration constant. The 0.5 power in the velocity term comes from the Schmidt number in mass transfer [7]. 3 S E V" "f-CONTACT LENGTH Figure 2.3: Schematic illustrating chip flow and continuity conditions [Kramer and Suh, 1980] For a compound AxByCz, the solubility is given by equation 2.4. AG — AG" - AG" - AG" -RT 1 + l + 1n . CA 8, =EXpH 4"” x A y 3 z C (x nx y try 2 2)]] Equation 2.4 (x + y + z)RT where A043}; = Free energy of formation of the tool material AG ‘" A = Excess free energy of solution of component A in the chip AG” ,3 = Excess free energy of solution of component B in the chip 14 AG” (r = Excess free energy of solution of component C in the chip R = Universal gas constant T = Absolute temperature in Kelvin Notes 1. a-Iron is assumed to be the phase into which the tool material dissolves even though the transformation to the y-phase occurs at 996K. 2. Regular solution behavior is assumed. Because A and B cannot be determined directly without experiments, the previously discussed models can predict only relative wear rates where relative wear rate is defined as a ratio of wear rates between a candidate coating material and a reference coating material. 2.3 Comprehensive Models for Crater and Flank Wear Many workers [10,15] have enunciated flank wear to be a case of pure abrasion. Hence flank wear can be modeled as a case of abrasion alone [equations 2.1, 2.2] and hence wear needs to be calculated as a wear volume for this purpose. Furthermore, It is widely believed [11,16] that flank wear rate is best expressed on a sliding distance basis in view of the fact that unequal cutting times can give a distorted picture of flank wear. It can be seen in Shaw [25] that b 2 tanB , = %— Equation 2.5 Where B, = Volume worn away on the relief face w = Wear land on the tool 0 = Relief angle of the tool. Hence the wear volume on the flank is proportional to the square of the wear land. Since the Rabinowicz model calculates wear on a sliding distance basis, the appropriate 15 [25] formula for calculating the abrasive wear rate on the flank derived from equations 2.1 and 2.2 is shown in equation 2.6. Flank volume wear per Sliding Distance = AK (Pam) )/(Pr") (3-body abrasion) = A(l/Pt) (2-body abrasion) Equations 2.6 where K is the factor from Equations 2.1. Experimentally obtained wear land values have to be expressed as (VB)2/Sliding-Distancel [Figure 1.1] to correspond to the abrasive wear calculation and A is a constant to be obtained from experiments. For modeling crater wear, one has to account for both mechanisms [13]. Wear rate has to be expressed on a time basis since the dissolution wear rate is not expressible on a sliding distance basis. Therefore [7,8] Crater Wear Rate = [AVK (P,‘"'” )/(P,“) + BMV0'5 chgycz ]. Equation 2.7 It is therein assumed that the abrasive and dissolution volume wear rates are directly proportional to the crater depth and hence experimental wear rate is simply KT/cutting time [Figure 1.1]. 2.4 Accounting the variation of hardness with temperature Most importantly, one has to consider the thermal softening of . the tool and the abrasive while machining. A suitable relation has to be evolved to account for the softening of the materials concerned. Kramer and Kwon [7] used an exponential function of the form of equation 2.8, H(T) = Hoe’” Equation 2.8 ' Sliding-distance (m) in a turning operation is cutting velocity (m/min) x cutting time (mins) 16 where H(T) was the hardness of a material at a given temperature 7°C. H0 and a are constants obtained form a curve fitting process on the empirical data. In the present analysis, the same form is adopted along and developed along with the other thermophysical properties in the computer algorithm to be described later on. 2.5 Comparison with Taylor’s model for tool wear F.W.Taylor was perhaps the first person to recognize the importance of cutting tool materials to modern civilization. Often proclaimed as the co-inventor of high-speed steel, his tool life criterion has been in use ever since he came up with it. Stated simply as in equation 2.9, it relates various process parameters and the tool life in an heavily empirical form. W"a"’ f ' = C Equation 2.9 where V is the cutting velocity, T is the tool life, a is the depth of cut, f is the feed and 1, m, n and C are constants obtained from a series of machining tests performed as specified in the ISO standard 3685-1977(E) [1]. However, simple as it may be, from the standpoint of adoption in the industry, it cannot be sustained indefinitely since a large number of work materials such as metal matrix composites and multi-layered coated tool materials, with a equally enormous number of geometries, are now becoming available. Not only does tool life has to be known apriori for given process parameters, the predictability of the machining process is equally important because of the close tolerances and stipulations being made nowadays on the surface finish of the component. Chapter 3 CUTTING TOOL TEMPERATURES Temperatures in metal removal processes have long been of interest to many researchers. There are essentially three temperatures to be concerned with in the turning process [Figure 3.1], (1) the shear plane temperature, which represents the bulk of heat generation in the chip formation process, (2) the chip-tool interface temperature, which is influenced by a number of factors including the secondary shear zone and (3) the flank temperature or the work-tool interface temperature, which is considerably lower than the other two. Flank temperature is important for its influence on the wear mechanisms prevalent at the flank in addition to its minor contribution to the heat generated in the chip-formation process. This chapter elaborates on the theory for the temperature measurement in the experiments conducted. chip-tool interface workpiece work-tool interface shear plane Figure 3.1: Sources of heat generation in a cutting tool 17 18 3.1 Inverse Problem of measuring interface temperatures Cutting tool temperatures are usually measured using a chip-tool thermocouple technique [11]. This technique essentially gives the area weighted mean of the temperatures at the chip-tool interface, the work-tool interface and the shear plane. Other models developed recently such as the 1D ellipsoidal model of Yen and Wright [26] avoid the problem of inaccessibility of the interfaces and determine the cutting tool temperature as a whole field measurement and thus enable inverse estimation of interface temperatures. These involve measuring the cutting tool temperature away from the interface and adopting an inverse estimation of the chip-tool interface temperature. 3.2 1D ellipsoidal model of Yen and Wright[26] The model essentially assumes that a one-eighth ellipsoid [Figure 3.2] in the domain of the cutting tool represents an isothermal surface. The cutting tool is therein assumed to be a semi-infinite body. Heat generated at the shear plane and the chip-tool interface manifest themselves as heat input into the cutting tool in a defined elliptical area of contact on the rake face. This heat input is represented as a constant temperature in this elliptical area and used as the boundary condition for the heat diffusion equation. rake face chip-tool contact area (base ellipse) “an... (>0 . ,o .9"! a I“ ...... I ,- ........ 0" “a! I. ......... '''''''''' .......... . .. “v“.f ........... ..... ................ ........ Figure 3.2: Temperature distribution in a cutting tool for the inverse temperature estimation 19 For the steady state temperature distribution, one has to solve the Laplace equation in ellipsoidal coordinates. However, it can be simplified for the special case of the one- dimensional (l-D) steady-state problem, where the temperature distribution in the tool body is a function 0, only of the radial coordinate g as in equation 3.1[26]. d d0 . E£Rg 61—5) = 0 Equation 3.1 . T; -T.. wrth R¢=\/(a2+§)-(b2+§)-§ and o=T T R — co where r, is the radial coordinate in the 1-D ellipsoidal model [mm2] a,b are the parameters describing the base ellipse with a>b [mm] 0 is the relative steady state temperature TR is the steady-state chip-tool interface temperature [°C] Tg is the temperature at the location determined by 5, [°C], and To. is the ambient temperature [°C]. The boundary conditions specified are: - All other faces are insulated (heat convection = 0 or negligible). Therrnophysical properties of the tool material are constant. The tool is rigid and tool wear is negligible. Uniform temperature TR at the chip-tool interface. 9993”?” Temperature at infinity or the far field of the tool is T... With some mathematical manipulation, a function for 0 can be found as in equation 3.2. a .5 l 2K0") [Mkaz +x)-(b2 +x)-x 2 iFESin-{J— 72% '32—] =1+ [([02 422] 02 dx @(4‘)=1— Equation 3.2 20 where F (1; | or) is the complete elliptical integral of order or. K(m) is the complete elliptic integral of the first kind, and x is a variable of integration. The final form of the solution for the one-parameter case (a = b) appears as: - Equation 3.3 where T, is the steady-state cutting interface temperature T is the steady-state remotely measured rake face temperature T... is the temperature in the far field of the tool (Taken to be 25°C) x is the distance of the point of measurement from the origin of the axes, and a is the radius of the circular tool-chip contact area. While the model can be developed more generally for three dimensions, closed form solutions exist only for the 1D case with one, two or three parameters. The one-parameter scheme is extremely efficient and facilitates accurate computation of the interface temperature. Hence from knowledge of the temperature at a point located remote to the interface, i.e. at a distance x from the interface, the mean interface temperature at the interface can be estimated using equation 3.3. This is the form for the inverse estimation which was used in this study. The inverse estimation scheme is however sensitive to measurement error, Lin et al. [27], while implementing it for the infrared method, concluded that the inverse predictions for the heat conduction problem is an ill-posed problem and instabilities are 21 likely. However, they established that temperature measured in the IR system was in agreement with that predicted by the metallurgical approach at low cutting speeds without chip control. They had also used a square tool in their experiments. Hence small deviations from theory are likely in using the same scheme for tools with a clearance angle, as in this study. The cutting speeds for which the effectiveness of the scheme was tested was also comparatively higher than that used in this study. In a similar approach, Rall and Geidt [28] concluded that most of the energy in the the metal cutting process appears as heat that is transferred to the chip, even though the small proportion of heat transferred to the tool influenced tool life. 3.3 Oxley's method of obtaining the flank temperature The temperature difference between the chip-tool interface and the flank has been a much debated topic. In the literature survey presented before, one can see a number of opinions raised regarding this. In this study, the relation proposed by Oxley [29] was undertaken. Oxley’s conclusions were drawn from the isotherms of the infrared photographs taken by Boothroyd in an earlier work [30] where stresses and temperatures on the flank wear land were studied. Boothroyd’s work also concluded that large variations with temperature do not occur (<16°C) over the flank wear land and that temperature was higher away from the cutting edge. Stated simply, the tool work interface temperature is 0.82-0.95 times the mean chip- tool interface temperature (All temperatures being in Kelvin). An average value of 0.89 was used for this study. Slight deviations are possible due to the sharp HSS tools used in the Boothroyd’s tests [30]. Moreover, the higher thermal conductivity of the carbide tools [12] would tend to make this factor higher than that for HSS. In the absence of an exact dependency, this nevertheless suffices for this context. Chapter 4 EXPERIMENTAL WORK 4.1 Turning Experiments The experimental work was carried out at the premises of the Lansing Community College in Lansing, MI. The machine used was a Milltronics Manufacturing 20 HP medium sized lathe with provisions for infinitely variable speed and programmable feed and depth. It had a rigid tailstock, important for ensuring minimal chatter while turning long bars. The lathe was facile since the RPM could be controlled to keep the cutting speed constant. Figure 4.1 shows the photograph of the lathe with the experimental set- up. Figure 4.1: Photograph of the lathe with the experimental set-up (Inset: pyrometer end probe) 22 23 4.1.1 Cutting conditions and related issues Dry cutting experiments were performed at a constant feed of 0.3 56 mm/rev @.Ol4"/rev1 anl depth of 1.905 mm (0.075") while the cutting speeds selected were at 300, 500, 700 and 900 sfpm (between 90 to 275 rn/min). Cutting speeds were chosen after referring to the insert manufacturer’s recommendations. The feed and depth of cut were determined after a few trial runs with the inserts to give a consistent cut with minimal chatter, sparking and good surface finish while maximizing on cutting time to reach steady state cutting conditions. The machine was also programmed for the RPM rather than the cutting speed to keep the cutting velocity constant. A rigorous study of flank wear has to take into account this effect of the “cut-in” wear, which develops on the flank immediately after the commencement of cutting. Dearnley [9] had noted that flank wear rates estimated after longer cutting times were more representative than that after shorter cutting tests. In all cases at least a minute of cutting time was achieved to supercede the effect of this cut-in wear and ensuring that the tool reaches a steady state of wear. Exactly one cut was performed for every satisfactory data point. Data sheets consisting of the wear data and the process parameters were prepared for every cut and were compiled separately both as a hard copy and on the computer. The machining data (cutting speed) is shown in tables 4.6 and 4.7. 4.2 Method of Temperature Measurement As explained before, there are several ways to measure cutting tool temperatures. In this study, infrared pyrometry was adopted. Conventionally this method is suitable only for high temperature measurements such as that in ovens and other places. However, new detector materials have been developed and hence this efficient and quick way of non- contact temperature measurement has become accepted all over the world. This method was deemed more reliable than other methods of cutting tool temperature measurement such as embedding thermocouples because: - 1. Bonding of thermocouples to the cutting tool is always a problem, given the dynamic nature of the process. 24 2. Thermocouples will tend to thermally ‘load’ the cutting tool and may alter the temperature patterns on the tool. 3. The pyrometer technique is reputed for the fast response it possesses. 4. Since the volume of cutting tests to be performed was very large, a non-contact method saves a lot of time in interchanging and indexing the cutting tool. Coupling of fiber optics to infrared detectors represents the latest progress in the field of non-contact temperature measurement and control. Previously, fiber optics was excluded from consideration since they were made of glass and plastics, which are opaque to infrared radiation. This attachment can be beneficially used in cutting tool temperature measurement since it allows the probe to travel along with the cutting tool during the feed. 4.2.1 Principles of Infrared Pyrometry The underlying principle of this technique is that all matter emits electromagnetic radiation at a temperature above 0K proportional to the fourth power of the absolute temperature. A temperature dependent emissivity is the parameter that adds a dimension to the physics of the method. Though emissivity itself carries several meanings according to the circumstances, it represents the behavior of a real surface and can be generally defined as the ratio of the radiation emitted by the surface to the radiation emitted by a black body at the same temperature. 4.2.2 Implementation of the Infrared Pyrometer for the experiments The detector head of the pyrometer was designated 0S1513 General Purpose Sensor used in conjunction with Model 3026 Single Channel Thermal Monitor]. It possessed a response time of 10-msec and was calibrated for the temperature range of 84- 300°C. The end probe of the fiber optic cable was a glass tipped steel probe of 3" lengg (Not as shown in the photograph in figure 4.1) and the spot size was found to be 0.785 mm2 prior to the commencement of every cut. Accurate positioning of the probe was ' Courtesy OMEGA Vanzetti, Inc. 25 done using a backlight source. There was also a provision for emissivity correction in the system, though this was not used in this study. The calibration of the pyrometer was performed using the BB-4 black body calibration source’. The non-linearity was found to be less than 1%. The absolute accuracy of the pyrometer was found to be 3°C. Since the detector head was capable of measuring radiant energy in only one wavelength i.e. a single color pyrometer, it had to be corrected for the emissivity. Emissivity is always an important consideration in such non-contact temperature measurements especially while using these techniques for metallic surfaces, which are usually shiny. A high emissivity is desirable since it eliminates the adulterating effect of the background radiation. Among the coatings used for the testing, the TiN and TiCN inserts had a very shiny surface. In a concurrent study that was carried out, the emissivity of the TiN surface was found to be 0.101, which was dangerously low for temperature measurement. To circumvent these problems, a thin coating was black high temperature paint (Flat Black) was sprayed on the inserts prior to the tests. Proper drying of the paint was ensured since this could harm the measurement during the experiments. The emissivity of this paint, which was known to be 0.92 from OMEGA, was the value to which the monitor always set. It is recommended that in future, the dual color detector head be used since this can measure temperature independent of emissivity. 1 Courtesy OMEGA Inc. 26 BRAIDED CABLE END Prone. / ----------- 1 CHIP Baum-ta INSERT X Clamp _- Chip-Breaker <..- P...— __'-'- Hole for End Probe Test Insert / Chip-Tool Interface Figure 4.2: Illustration of the chip-breaker set-up over the insert 27 In order to prevent the interference of the chip in the field of view of the infrared probe and to exercise some control on the chip, a chip breaker was designed using one of the same inserts to hold the pyrometer in place. Figure 4.2 shows the schematic of the set- up shown to illustrate this point. It should be noted that there might be some disturbance in the temperature field due to this device. But in all cases it was found that it was sufficiently far away from the interface as shown in figure 4.2. The final assembly is as shown in figure 4.3. FmrmOmc . ’ Arnamarr l . ’ i l 5 f l Qiamlusra"1'"~ . , p _ / .' \. Figure 4.3: Final assembly of the pyrometer over the insert 4.3 Data Acquisition Temperature data was collected in an automated data collection system. This consisted of a data acquisition board, signal conditioning and software which was written for the purpose of collecting and storing the data in a spreadsheet. The timer of the 28 computer was used, enabling the on-line monitoring of the temperature. Though initially the software was written for acquiring data in 5 channels-2 temperatures and 3 cutting forces, only the channel for the pyrometer was ultimately used. The thermocouple underneath the insert and the dynamometer were not implemented. 4.4 Work Materials The work material consisted of steel rounds with the AISI designation 1018, 1045, 1065, 1070 and 1095. Being plain carbon steels, cementite constitutes the bulk of the inclusions in their microstructures. All the work materials for the experiments were procured from Alro Steel Corporation, Lansing, MI excepting the AISI 1065 bar, which was acquired from Timken as a compliment. It was learnt that the AISI 1065 steel was more commonly used as a plate rather than as a bar stock in the industry. 4.4.1 Dimensions of the bar-stock, cutting length and duration of cuts The work material used for the experiments were steel rounds commercially obtained fiom ALRO Steel, Inc., Lansing, MI. The bar stocks were nominally of diameters between 3” and 6” and length was about 2-1/2' initially. The 1065 bar alone was 2” in diameter to begin with. However, all the bars had been reduced to a diameter of about 1.5" at the end of the tests. The cutting length was different for different bars for stability reasons. As mentioned before, the shortest cutting time was 1 min and 3 sec. 4.4.2 Spherodizing-annealing of the steels The spherodize-annealing process transforms the cementite in the steel in to spheroids and brings the steel to a dead-soft condition, thereby removing any shape effect in the abrasion of the tool while machining. Those that were required to have a spherodized microstructure (1018, 1045, 1065 and 1095) were sent for heat treatment to Atmospheric Annealing, Inc., Lansing, MI. The sheer size of the bars excluded any possibility of annealing in a laboratory. This was also the reason due to which control over the grain size of the cementite in the steels was not possible. The as-received steels for the tests (1018, 1045 and 1070) were mainly in a hot-rolled, normalized condition. 29 The details of the annealing process for the AISI 1045 steel is shown below: - 1. 0-1000°F in 6 hrs, then hold for 4 hrs. .O‘SJ'PP’N Then cool to room temperature in air. 1000-12000°F in 5 hrs, then hold for 3 hrs. 1200-1310°F in 5 hrs, hold for 50 hrs. 1310-1200°F in 3 hrs. 1200-1150°F in 0.5 hrs. The process was similar for the other steels. 4.4.3 Microstructures, Composition and Hardness of the steels compositions of the bars are as shown in table 4.1 and 4.2. Table 4.1: Composition of the hot-rolled steels used in the testing (All in wt%) C p 8 SI 1N1 Cr Mo Cu Sn AI V The hardness and the composition of the steels are examined in this section. The T1 1018 1045 1070 0.208 0.476 0.684 0.702 0.744 0.780 0.015 0.011 0.013 0.026 0.037 0.024 0.212; 0.069 0.273 1 0.215111 0.042 1 0.051 0.133 0.077 0.165 0.018 0.015 0.016 0.258 0.109 0.046 0.011 0.004 0.006 0.020 0.037 0.020 0.003 0.004 0.000 0.001 0.000 0.000 0.004 0.005 0.004 Table 4.2: Composition of the spherodized steels used in the testing (All in wt%) C 4- p S 1018 1045 1065 1095 0.160 0.476 0.640 0.887 0.828 0.744 0.800 1 .024 0.010 0.011 0.014 0.019 ’ 0.028 0.037 0.010 0.025 31 "0"."1‘03’ 0.273 0.280 0.309 0.051 0.070 0.145 ""6361’4‘ Cr Mo Cu Al V ‘— Tl ”01079 0.077 0.150 0.316 0.011 0.015 0.020 0.141 "0.053 0.109 0.130 0.156 0.004 0.004 0.009 0.007 0.020 0.037 0.024 0.004 0.001 0.004 0.002 0.170 0.000 0.000 0.000 0.000 0.001 0.005 0.002 . 0.004 Figure 4.4 shows the photomicrographs of the microstructures. As one can see, the as-received steels showed a very pearlitic structure. For the spherodized steels, this lamellar cementite were transformed into spheroids by the spherodizing process. It is also evident that the cementite in the spherodized steels was not of a uniform size and this could have a role to play in the tool wear process. In particular, the 1095 steel showed a 30 very small cementite size, which was believed to be because it was vanadium killed [Table 4.2]. Figure 4.5 shows the hardness of the steels in various states. 1018(as-received) 1045 (as-received) 1070(as-received) 1018 (spherodized microstructure) 1045 (spherodized microstructure) 1065 (spherodized microstructure) 1095 (spherodized microstructure) Figure 4.4: Microstructures of the steels used in the testing 31 m: moron steels Dem 240 “WWW mmsnnoruumtaoaeu .BdoreSphnrodiziro 3 . . . 1 7 ‘ ; " uAiterSpharodizim ‘ Dachau [M a‘. o ' quuvmvm I 1018 1045 1070 AISI Dealngatlon of Steel 1018 1045 1065 1095 AISI Stool Damnation Figure 4.5: Hardness of the steels used in the machining tests 4.5 Inserts used in the experiments The inserts for the experiments were procured on request from Kennametal, Inc., Latrobe, PA. The turning tool holder was also procured fi'om them. Adequate number of spare parts for the tool holder were also purchased for exigencies during the experiment. 4.5.1 Insert Geometry and ISO/ANSI designation The geometry of the inserts had the ISO designation SPGN 19 04 12. The integral chip breaker was intentionally avoided (l) to simplify modeling of tool temperatures, (2) to make an accurate measurement using the pyrometer and (3) to apply the 1-D ellipsoidal model. A large nose radius was chosen to avoid catastrophic failure of the insert and a large IC (Inner Circle diameter) was chosen to suitably place the pyrometer on top of the insert. It was felt that a proper choice of the nose radius is important since this could bring about chatter problems and difficulties in flank wear measurement if chosen improperly. It should also be mentioned that the depth of cut should be at least as high as the nose radius to have a suitable chip curl. There were also many opinions that the insert style was one of milling than turning. The inserts were used with a standard CSRPR 856D tool holder. The assembled tool signature of the insert, measured with the toolrnaker’s microscope is shown in table 4.3. Figure 4.6 illustrates the various geometric parameters. 32 Table 4-3 Cutting Tool Signature I Parameter Back Rake Side Rake End Relief Side Relief —End Cutting Side Cutting Nose Angle Angle Angle Angle Edge Angle Edge Angle Radius Symbol a. at, 0C 0, cc c, r “We 0° 4°42' 4°42' 0° 15° 15° 3/64" {\ ISO specification PGN 19 0412 I Lnose radius :12 mm thickness 5 : 4.76 mm length i, inner circle 10: 19.05 mm insert type N; without hole, without chipbreaker tolerance class G relief angle P (beta): 11 degree shape 3: square, nose angle 90 degree Figure 4.6: Illustration of the various geometric details of the cutting tool 33 4.5.2 Grade of substrate used The substrate was a K4202 grade, which consists mainly of WC and a C0 binder has a slight alloying of TiC and TaC. The American standard for the same is C6-C7. The details are shown in table 4.4 [31]. Thermophysical properties of the grade collected from Santlranarn [32] are shown in table 4.5. It has a medium binder content and a large grain size meant for general purpose steel machining. It is also renowned for the right balance of wear resistance and toughness. Table 4.4: Properties of the substrate K420 ISO Specification Nominal Specifications P25-P35 Porosity A04-B00-C00 M25-M30 WC (Gsum) l-8um Nominal Composition — wt% HRA 91.2 Co 8.5 Hc(0e) 140 TaC 10.9 Dens. (g/cc) 12.65 TiC 7.4 TRS (Mpa) 2170 WC 73.2 Table 4.5: Thermophysical property data of substrates Composition Grain Hardness CTE —Thermal Density g/mmz TRS size HRA um/m°C Conductivity MPa W/m.K 94WC-6Co Fine 92.5-93.1 5.9 108 15.0 1790 94WC-6Co Medium 91.7-92.2 5.4 100 15.0 2000 90WC-10Co Fine 90.7-91.3 6.0 80 14.6 3100 72WC-8TiC- Medium 907-91 .5 6.8 50 12.6 1 720 1 1.5TaC- 8.5Co 2 Courtesy Kennametal, Inc. 34 4.5.3 Coating Materials and related details The TiN coating was commercially available from Kennametal as KC7102. It consisted of the K420 substrate with a 411m coating of PVD TiN over it. The two other coatings TiCN and A1203 were custom made for this study as follows. Uncoated inserts were sent to Balzers, Inc, Lansing, MI, where a 3.5pm coating of TiCN was performed. The PVD process used for this purpose was Reactive Ion Plating. Uncoated inserts were also sent to Valenite, Inc., Troy, MI, where a 311m CVD coating of A1203 was performed. Due to the well-known problems of the A1203 coating adhering to carbide substrate [32], a 111m intermediary layer of TiN was deposited between the A1203 and the carbide substrate. The nature of the CVD process used was not disclosed. The hardness data of the coatings was also collected from the literature and a variety of sources to be elaborated later on. The TiN coating was a golden colored coating whereas the TiCN had a bluish gray color and the A1203 coating had a dark black color. Also the TiN and TiCN coatings were shiny while the A1203 coating was rough and dull. This great difference in colors and texture persuaded the thin black coating in the temperature measurement as explained before. 4.5.4 Designation for the identification of the inserts Given the huge volume of tests that were performed, it was imperative that a proper designation system for the inserts be evolved. This was done and the same designation used for the data sheets, the temperature records and the crater photographs stored on the computer. Figure 4.7 illustrates a sample naming scheme used. Tables 4.6 and 4.7 Show the process parameters of the cuts with reference to this naming scheme. In all 84 cuts were performed, of which, two were invalid. 2 Courtesy Kennametal Inc. 35 xng1045un-li1 I > Cut Number of the edge ) Edge Number p Unspherodized/Spherodized > AISI steel designation » Insert alphabet or name > Coating TiN - n TiCN - 0 Ale3 - 0 , Grade of carbide used x - K420 Figure 4.7: Sample-naming of the TiN insert used on AISI 1045 (spherodized) for the low speed cut. 36 2...: 1.38.22. .28 4.2333. 3.38 3.8.8.. 3.5 2.3339. .2: 2.38.22. 83. 2.3332 8.8. 2.3889. 2.8. 33339. 6.2 6.2 3.8” 1.38:3 .28 1.3332. a... . a «238...: 3.8a 2.3332. 2...: 2.8.3.. 8.3. 2.4332. 3.8. 138:3 8.8. .2332. 2o: 5: .28 1.38:5. 3.8m 1.33.»: we. . a 2.38:5. 3.38 2.33.? 8.3. 2.38. .5 8.3. 2.3325. 3.3. 338:5. 8.8. 133.9: 2: 2338. 2; 2332 2.5 1.38.80. «38 «-22262 2.2... 2332 .6 6.2 s... so 3.: 328 $2228.. 3...: 3.328.. 3.3. 22228.. 322 .2322; 8.8. .2222... 6.2 m2.... 2 . .3 4.28.8. 3.48 25:28.. 82.3 «.2383: .33 4.2.8.8.. «:2 22328.. 8.3. "2228.. 8.8. 138.80. ~22 .2228x 5: 2o: 5.8 1.3825. 8...: 138:5. as: 13825 Son 133...: ~32 3.32.5. 8.3. 02235. 2.8. 13825. 8.8 .2235. 2:. 2232 2: 2222 .5525 82.5.3 .5525 8.65.3 68%. 9.25 3.an 9.2.8 .3an 0.33 :3... Sam 225 arses. 9.230 3.an new use. wage. £86 3333...... 05 .8 £808.53. 808k— 66 03a... 37 2.35 1.3332. 8.8 1.3332. 38. 8.3332. man 2 8.. 35.8.... .3 a... 6.2 8.38 1.3332. 8.88 2.3332. 8.8. $533.32. 8.... ....3332. 2o: 8.8 $833.55. 8.8 $833.55. 8.3. «5:33.85. 8.8 7533.5... 23 £525.33. 8.nh~ I 4 . ||-.. --.-.. .. . . ..a V'goho_=°x om.th ......yi....................... .. .. . .. . ......C... ........I-I|.Ii...!..........tb..y - uvgw~°_—°x 833 $888.52. 8...: $822.2. 8.8. $558.52. 8.3. $822.2. 8.8 .-558...2. 8.8 75.22.... .03... 5.2 8.8 755832. 838 5822.2. 8...: $558.52. 833 $822.2. 8.8. $558.52. 8.3. $822.2. 8.8 7558.5... 8.8 .8222. 20... 20: 2.58 7888.5... 8.38 5558...... 8...: $558.5... 853 $55.25.. 8.8. $5.825... 8.3. $822.5. 8.8 7558...... 8.8 7558...... 2: 32588. 2: 3.55.8.2 A58)... ASE—5 82.3 3.2.6 .5553 3.58 522.3 .83 8.5.5.3 8...... 82.3 3.5.0 .5853 3.58 3.2.5.. .83 8.5.5.80 82.. wage. ...—08m Egunams. o... .5. flaogm 8.89... ....v 035... 38 4.6 Area Measurement Photographs of the crater were taken with a LECO SZH stereo microscope with a LECO 2001 image analyzer. Samples of the photographs taken are shown later in figure 5.4. Measurement of chip-tool contact area, which is a vital input in the inverse temperature estimation, posed a great challenge. After attempting a number of methods, it was found that it was most accurately measured using an image processing software. However, the spatial calibration was a critical issue. Figure 4.7 shows the methodology used in this regard. The diagonal formed by the ends of the nose radius was the easiest and the most accurate dimension, available right on the insert specimen. It also avoids any parallax error, if any. The image processing software also enabled adjustment of the contrast of the image. This was important since precise demarcation of the chip-tool contact area was sometimes absent. Nose Radius ——\ __,..--""i .6838 mm Insert Figure 4.8: Illustration of the method for the area calibration 4.7 Tool Wear Measurement : set-up for Flank and Crater Wear Flank wear was measured using the Mitutoyo TM-505 toolmaker’s Microscope at a magnification of 200. The microscope was equipped with digimatic heads enabling measurements accurate up to 1 um. The set-up is shown in figure 4.8. Crater wear >5 um was measured on the stage and the optics of the microscope at a magnification of 30, with the Fowler D1040 Digital Test Indicator. The accuracy on this was lum. The calibration was checked using one-tenth thousandths gage blocks. For very shallow craters, the Sloan 39 Dektak IIA surface profilometer in the premises of the Physics Department of M.S.U. was used. A sample of the profiles obtained is shown in figure 4.9. Figure 4.1: Set-up on the TMM for measuring flank and crater wear 10:2 SCAN: 3,000 mm VERT: 0 A 17:07 02-26-98 SPEED: HIGH HORIZ: 1,930 um 80,000 75,000 70,000 65,000 60,000 0 500 1,500 2,500 A CUR: 01,301 A Q 1,052 um M CUR: 01,301 A Q 3,988 pm SLOAN DBKTAK 11 Figure 4.2: Sample of crater profiles (Cut-ID xoa1018sp-2) Chapter 5 RESULTS AND DISCUSSIONS 5.1 Results of the Computer Algorithm The results of the computer algorithm [Appendix A] yield a general ranking of the three candidate coatings fi'om the hardness and chemical properties. A coating with a greater hardness has a better abrasion resistance and a coating with a lower free energy of formation has a greater resistance to chemical dissolution. Though cementite is softer than the coating materials at all temperatures, the variation of the abrasive wear-rate with temperatures, as given by the theoretical formulae in Chapter 2 was more keenly seen. Interestingly, 3-body abrasion showed a decreasing rate with temperature. This is justified since even in the work of Rabinowicz [33,24], a rapid fall in wear-rate was observed with decreasing hardness of the abrasive. Two-body abrasion, however, shows an increasing rate with temperature, simply because the hardness of the abrasive is assumed to be infinite at all temperatures and only softening of the tool material is taken into account. From a practical standpoint, the validity of this assumption is indeed questionable. Dissolution, being inherently a thermally activated mechanism, increases steeply with temperature. The exact methodology of the numerical computation is elaborated in Appendix A. Figures 5.1 and 5.2 show the output of the computer program for the three coatings in absolute terms. 40 41 Theoretical 3-Body Abrasive Wear Rate 1.E-04 l" I i _.°_..TiN . . .g. . . TICN 1E'06 +Alum'na Volume Wear Rate (Arbitrary Units) 1 E-OB . El ..... . , a ....... B- """"" 13 1.E-10 4 . 1 700 900 1 1 00 1 300 Temperature (’C) Theoretical 2-Body Abrasive Wear Rate 6.E-03 l Volume Wear Rate (Arbitrary Units) 700 900 1 100 1300 Temperature (°C) Figure 5.1: Charts illustrating the ranking of the coatings for the two abrasive wear mechanisms. 42 5.1.1 Abrasive Wear It can be seen in figures 5.1 and 5.2 that wear-rate decreases with temperature for the 3-body abrasive wear model whereas it increases with temperature for the 2-body wear model. A curve fitting of the form Ae‘“T was performed for figure 5.1 for prediction of wear with the experimental results. The final equations evolved for the 6 cases are shown in table 5.1. Table 5.1: List of Wear-rate Equations for the calibration purpose (T in °C, w.r. stands for wear per sliding distance) Coating Compound Three-body Wear Equation Two-Body Wear Equation TiN w.r. = 7.5506E-04e'9'3077E’03T :w.r. = 3.91735-04e‘596254’“ TiCN w.r. = 3.629515061-2”0°89“3T Ew.r. = 1.8166E-04e2'°°”E'°3T i 1 N20: w.r. = 1.374230465751550” §w.r. = 30735150413.“74°“3T i 5.1.2 Dissolution Wear As can be seen from figure 5.2, dissolution wear-rate increases with temperature. Also, A1203 is considerably more inert that the other two coatings. This corresponds to the common experience of using A1203 coatings for finish machining since higher temperatures are encountered due to the high cutting speeds employed for finishing. The curves were then fitted in the form Ax“, where x is the temperature in °K. The reason for using this form was the steep drop in dissolution wear with decreasing temperature. The final equations are shown in table 5.2. 43 Table 5.2 Curve fitted relations for the dissolution wear of the coating materials (x is tern rature in °K, w.r. stands for wear r unit time) P9 pe Coating Dissolution Wear Equation TiN w.r. = 1.1155815-54x1-6234N1 ncn w.r. = 185755431252"E+O1 N20: w.r. = 8.1920E-96x2'82°15*°‘ 1.5E-02 F— 1 "C W :3 g l -9— TICN 5 g 1.0E-02 ll A— Alumina <> 2 5 m ‘ a , 2 5 ° 2 8 s. 2 O > HS 1300 Temperature (°C) Figure 5.2 Figure illustrating the ranking of the coatings for the dissolution wear mechanism. 44 5.2 Temperature trends with cutting speed A sample of the temperature profiles for the cuts on spherodized steels, TiN coated K420 while machining spherodized AISI 1095 steel, is shown in figure 5.3. One can note an increasing trend with cutting speed. The curve fitting, to obtain the steady state cutting temperature, is explained in Appendix D is shown in the same graph. 180 _ 160. 2m" -' i 140 '1 « 1 "5" mm *1 l I nIfl-Qg-W 1W . r :v 1“ I. I” - / . i -c' 1' r _, . ' 120 ' " ‘ ' I .l 1 fl . IJ : ; ' v 1 v 1 V Temperature (°C) 0 20000 40000 60000 80000 100000 Tina (ms) Figure 5.3 Temperature records of cuts on AISI 1095 steel with TiN coated K420 carbide (Curve fitting process shown alongside) Although the pyrometer technique was easy to conceive and implement, the accuracy in measurement had to be validated with the data obtained in the past. The chip- tool interface temperature measurements compared favorably with the results obtained by Subramanian et al. [20] who employed the same work material, similar cutting conditions and grade of carbide. They had estimated the interface temperatures based on the cutting forces of Boothroyd’s model [30]. The flank temperatures observed in the present experiments showed an increasing trend with the cutting speed, except in a few cases [Appendix C]. Complications due to abnormally large contact areas may have contributed to the high temperatures observed in these low speed cuts. Another possible reason could 45 be that the peak temperature was far removed from the cutting edge thereby decreasing the apparent distance of the pyrometer to the interface. The inaccuracies in the inverse temperature estimation schemes such as the 1-D ellipsoidal model is fortunately low since the variation in the temperature field in the chip-tool interface is mitigated at the far field by the dampening of higher spatial temperature frequencies by the tool body [26]. Though the curve fitted peak temperature based on pyrometer readings [See Appendix D] generally increased with cutting speed as shown in figure 5.3, there were many instances where the pyrometer did record lower temperatures at higher speeds. The reason suggested is that the interface temperature in the inverse estimation was a function of both the pyrometer temperature and the contact area (which showed a decreasing trend as explained in the following section). This is direct evidence of the presence of steep temperature gradients in the cutting tool during cutting since this implies that the temperature of the area of heat input into the tool is far higher than the temperature outside this region. From the argument of the ellipsoidal model, this stands justified since, the volume for heat removal is higher farther away from the tool [figure 3.2], and hence represents a greater capacitance for storage of heat. The trends of the rake face temperature with cutting speed for the as-received (normalized) steels were not as uniform as that obtained with the spherodized steels [Appendix C]. The pearlitic structure of the steel, which represents a lower mean-ferrite path [25], could have made the flow characteristics of the steel substantially different from that of the spherodized counterpart. The undetermined pre-work in the steel could have also influenced the chip-tool interface temperature. 5.3 Contact Area Trends A significant finding in the results [Appendix C] was that the chip tool contact area showed a decreasing trend with cutting speed in both the spherodized and unspherodized cases. Subramaniam et al. [20] observed an increase in the crater area with cutting speed. The difference between the chip tool contact area and the crater area is that while 46 the chip-tool contact area corresponds to the region of heat input into the tool and the area of partial or complete seizure [40], the crater area corresponds to the amount of coating and substrate material carried away by the chip. The contact length was similar among the various coatings, which was also observed by Dearnley [9]. As is evident from figure 5.4, crater wear commences sooner at a higher cutting speed, possibly due to the chemical inertness and the resistance to surface traction offered by the coating material at lower speeds. It can also be noted that the alumina coating did not wear as much as the other coatings at the same cutting speed Considering the remarkable chemical inertness of A1203 [7], this elucidates the thermo-chemical component of crater wear. TiN TiCN A1203 Figure 5.4: Photographs of K420 carbide coated with TiN, TiCN and A1203 respectively alter machining spherodized A181 1045 steel at increasing cutting speeds at f = 0.356 mm/rev and doc = 1.905 mm. 47 Among the various work materials investigated, the chip-tool contact area showed a slightly decreasing trend with carbon content of the steel. The contact area is conjectured to be a strong function of the fracture toughness [Appendix C] of the work material, since this property governs the deformability at the shear zone. 5.4 Flank Wear Flank wear has widely been recognized as the most appropriate criterion for tool life [11]. Chubb and Billingham [15] found that the removal of the coating in the region of the flank could accelerate the rate of wear. Hence, the resistance of a coating in the progression of flank wear deserves greater attention. Dearnley [9] and Cho and Komvopoulos [10] concluded that the WC phase in the substrate is prone to dissolution into or or y phases of steel and dissolution wear supercedes any other mechanism of wear for the case of uncoated carbide grades without any TiC or TaC alloying [13]. In the present work, as expected, the flank wear-rates increased with the cementite content of the steels. As noted by Kim and Durham [17], the region of the flank wear could be divided into several zones of damage which includes the zone reminiscent of superficial plastic deformation and plowing by small carbide grains [10]. In this study, flank wear was measured at the region of wear land exhibiting uniformity in wear pattern. Ramalingam and Wright [2] had noted that alloy chemistry does not satisfactorily explain the machinability of steel since nominally identical alloys yield different machinabilities in different heats. Thus, it has been corroborated herein that the constitution of the steel in terms of its hard and soft phases is responsible for flank wear and hence its machinability. Abrasion by other means (such as the wear debris generated by the tool) cannot give rise to a uniform wear pattern since this form of abrasion is stochastic in nature. The effect of competition between the various mechanisms in coated tools is the problem at large and requires careful introspection. It was seen in the present work that the A1203 coated inserts showed scouring marks uniform in length from the cutting edge 48 [Figure 5.5] while the other coated inserts showed scouring marks but with a shiny band where the carbide substrate was exposed. It can also be seen in Appendix C that TiCN coating showed the lowest wear-rate. In the context of coated tools, it is common experience that coated tools exhibit a lower crater wear-rate due to the impedance to solid solution formation offered by the coating. Given these facts, the superior chemical inertness of A1203 [7] and the higher hot hardness of TiCN [31], it can be concluded that temperatures prevalent at the flank do allow preferential dissolution of WC to take place but the progression of flank wear is resisted by the mechanical superiority of the coating. This resistance is a property greatly determined by the hardness of the coating at the flank temperature. It should be noted that the preceding arguments would be true only if abrasion is the predominant mode of mechanical wear. This may not hold if more complex phenomena such as thermal cracking, mechanical fatigue cracking, chipping or fi'acture start adding to the damage of the cutting tool. 49 100 m/min 270 m/min TiN 00 .E ‘3 Z o ‘11-: Q o {-1 c: .2 H 8 .5: D M NW“ wa~u,.s.r~' 's..,..~....,__. . C A.“ Figure 5.5: SEM micrographs (800x) of the wom flank surfaces of the coated cutting tools after machining spherodized A181 1045 steel for low and high cutting speeds. 50 In the overall analysis, the TiCN coating showed the highest resistance to flank wear. A decreasing trend in the wear-rate with flank temperature was noted in the case of spherodized steels. This can be attributed mainly to a higher softening rate of the abrasive particles. Suh [19] mentions that Fe3C is capable of dissociation at high temperatures, vis-a—vis oxide and nitride inclusions. This is also a possible reason for the decrease in wear-rate with temperature, though not quantitatively ascertained. In particular, the alumina coating exhibited the greatest sensitivity to flank wear due to a higher softening rate [7] reflecting the contrasting properties of the A1203 and the TiCN coatings. While Brun et al. [4, Section 1.2 (paragraphs 2,3,4)] had observed excessive wear-rates while machining silica reinforced aluminum and the same was the case with Ramalingam and Wright [2], wear was tolerable in the machining of plain carbon steels. In view of the fact that SiC, SiOz and A1203 are more refractory materials than cementite, as can be seen from their hot hardness [7], they can inflict a greater damage on the tool than cementite. One can therefore generally conclude that inclusions and reinforcements in a material should be chosen so that they strengthen the material but do not degrade machinability at higher temperatures encountered in high speed machining. The final part of the analysis is an attempt to calibrate the comprehensive wear equations for the three coatings. Figure 5.6 shows the relevant graphs for the flank wear trends for three coatings. As can be seen from the figure 5.6, there is a convincingly decreasing trend in the spherodized steels and a moderately increasing trend with the unspherodized steels. The increasing trend in the case of the latter can be mainly attributed to the fact that the cementite particles are firmly rooted in the ferrite matrix. This supports modeling of the former with the three-body model and the latter with the two-body wear model. The decreasing trend should not give the notion that flank wear decreases with cutting temperature, rather, the rate of damage on the flank decreases with cutting temperature. The trends are not clearly evident with TiCN since the hard coating [31] was impervious to flank wear, due to the inadequate duration of the cuts. This testifies to the fact that abrasion is the rate-controlling mechanism on the flank. TiN o . 1018(sp) 30 , :1 1045(sp) 25 x Ag--- 1065(sp) A 20. E _._.___10958 a 15 . , x ( p) 5 i X C] A“A 7.x oi 33"%"9<>'° 700 900 1100 1300 Flank Temperature ('C) TiCN ..- 0 __1018(sp) 30 _ --- 1045 s 25 1:1 ( P) E 20 l A 1m5(59) : 15 1 . x» 1095~ 36:11.1 0 + I I 8 SE! 1 800 1150 1500 Temperature (°C) 1018(un) TiCN "1045(un) l 6 ‘ IE “A_1070(un) 4 A“. 1: 2 EA] 0 __ .' . 0 9 ° 334 o 5 800 1150 1500 Temperature (°C) 0 1018(un) ; A|203 -_-{}-_1045(un) | 5 +1070(un) E 4 _ 2 .' A?” """ A El 1 A: 13.----- 3 El 0 l 99 . . 800 1150 1500 Temperature (°C) 54 5.6 Calibration With use of the equations in tables 5.1 and 5.2 one can obtain a calibration for the wear-rates based on equations 2.6 and 2.7 and hence values for the constants A and B. As a recapitulation, the value for 'A' has to be obtained for flank wear, since it is dominated only by abrasion and both 'A' and 'B' for crater wear since both mechanisms operate. Further, for flank wear, it should be remembered that the three-body wear mechanism is assumed to operate in the machining of spherodized steels whereas the two body mechanism is assumed to operate in the machining of unspherodized steels. Furthermore, in the case of abrasion on the crater only the three-body mechanism is assumed to operate, since temperatures are excessively high for the abrasive to retain any hardness. The details of the governing relations of the wear mechanisms were explained in Chapter 2. As an example, calibrations for A1203 for flank wear and crater wear while machining spherodized steels is illustrated in figure 5.8. This was performed on the software SigmaPlot®. The values are listed in Appendix C. Figure 5.9 shows the plot of the constants 'A' and 'B' obtained from the calibration process. There is a general agreement in the trends of the models and that of the experimental results, however, it was not impressive for the as-received steels probably because of the pre-work in the steels and presence of distributed hard spots, which injure the tool in an arbitrary fashion. Since ‘A’ and ‘B’ are quite different in their order of magnitude, they are plotted on a logarithmic scale in figure 5.9 for visualization purposes. It can also be observed in figure 5.9 that the values for individual 'A's and 'B's are within the same order of magnitude. This is promising since this substantiates the correctness of the modeling, despite the fact that only one cut was taken for each machining condition. Volume Flank Wear per Sliding Diatance(‘m1Im) Wear Rate ( ’mlmin) 55 Flank Wear Calibration for AIzO, 30 .. o 1018(sp) 25 _, t predicted 1018(sp) \\ x. A 1045(sp) 2° ‘\ _ _ _ _ predicted 1045(sp) 15 ° A35." + :1: 1065(sp) 10 7 + s :; ... ‘ ....... predicted 1065(sp) . O ... 1095(sp) 5 A _ . __ . _ predicted 1095(sp) 0 a . . 5 700 850 1000 1150 1300 Flank Tem perature (°C) C rater Wear Calibration for A1203 6 1 0 1018(sp) 5 + predicted 1018(sp) ‘ A 1045(sp) 4 1 ......... predicted 1045(sp) 3 x 1065(sp) _A- '+ predicted 1065(sp) 2 +1, i + 1095(sp) . X A .. . predicted 1095(sp) o 7,_ _ ..-“ , . . 800 1000 1200 1400 Crater Tem perature ('C) Figure 5.8: Plot of the predicted and the experimental wear data for A1203 while machining spherodized steels Flank Wear Calibration for spherodized eteele 30 - 25 atom-2) % rout-p) 3 20 roost-p) E 13100500) 15 g 10 '6 " 5 2 l O , 10000000 [ , # ... 1000000 3 ‘3 100000 ‘ % a 10000 1‘ > g 1000. In a: <" _e 100 s 10 1 . 10000000 . 7 Hi ,A,,, 7,_ 1000000 8 % 100000 1‘ 3 D 10000 Te :1 ' i; g 1000 <- § 100 .j 5 \ Crater Wear Calibration for epherodized eteele 56 A (Arbitrary Units Flank Wear Calibration for Unepherodized Steele 4000 1 3000 amount) .10451011) I 1 070(un) O TIN TiCN Al203 D 1018(sp) A value I 1018(sp) B value I 1045(sp) A value D1045(sp) B value I 1065(sp) A value I 1065(sp) B value I 1095(sp) A value 1095(sp) B value D1018(un) A value D1045(un) 8 value Crater Wear Calibration for unepherodized eteele TiCN Al203 I 1018(un) B value I 1045(un) A value I 1070(un) A value I 1070(un) B value Figure 5.9: Plots of the constants obtained from the calibration process. (sp - spherodized steel, un — unspherodized steel) Chapter 6 SUMMARY AND CONCLUSIONS The following summarizes the work performed: - Wear of the flank and the crater of coated carbide tooling while machining plain carbon steels was investigated. Wear models in tribology and the literature were adopted in modeling and studying the phenomena. An attempt was made to relate tool wear and cutting temperatures in a quantitative fashion. The infrared pyrometry technique was utilized in the measurement of rake face temperature. The 1D ellipsoidal scheme of Yen and Wright [44] was used in the inverse estimation of the chip-tool interface temperature. Oxley's recipe for the flank temperature was adopted. The dependence of interface temperatures was on the process parameters was studied. The infrared pyrometry technique proved to be a versatile and easy method enabling easy measurement of the rake face temperatures of the tool, but relied heavily on the accuracy of the inverse estimation schemes for its accuracy in prediction. The related problem of the interfering chips was efi‘ectively dealt with. The three-body and two-body wear models were used for describing abrasion of the flank surface of spherodized and unspherodized steels respectively whereas the dissolution and the three-body wear models were used in obtaining a comprehensive description of crater wear in both steels. Recommendations for further developments in 57 58 tool materials such as multi-layered coatings and work materials include having an intermediary layer resistant to dissolution and an outer layer resistant to abrasion. The criteria for the predictability of tool wear have been outlined clearly. Reinforcements in a work material should be chosen so that they don’t degrade the machinability of a material. Among the major findings in this study are that steep gradients in temperature are present on the rake face of the cutting tool. Chip-tool contact area decreases with cutting speed. Contact area is proportional to the fracture toughness [Appendix C]. Flank and crater wear were found to increase with the cementite content in the steel. Spherodized steels showed a decreasing flank wear per sliding distance with temperature whereas hot- rolled, unspherodized steels showed a moderately increasing trend. Both types of steels showed an increasing rate of crater wear with temperature. Modeling of tool temperatures, which was undertaken, showed that there were steep temperature gradients in the cutting tool and temperatures depended on the area of heat input into the tool. The appropriate conclusions drawn include affirming that abrasion was the rate- controlling wear mechanism in the wear of coated carbide tools while machining plain carbon steels, though preferential dissolution of the carbide substrate is also possible. Crater wear involved both abrasion and dissolution. The wear models used in tribology proved to be reasonably accurate in predicting wear rates on both the crater and the flank. APPENDICES Appendix A COMPUTER ALGORITHM A.l The main part of mainprog.f character coating*8,elea(6)*2,compound*8 OOOOOOOOOO (10 30 31 52 53 68 69 70 ‘00000000 integer xa(6) 'Coating' IS THE INPUT VARIABLE FOR THE COATING, 'elea' IS THE GLOBAL ARRAY CONTAINING THE ELEMENTS OF THE COATING, READ FROM THE FILE COMP1.DAT THERE ARE FIVE FILES IN ALL. MAINPROG.F WHICH IS THE MAIN PROGRAM FILE, MECH1.DAT, WHICH CONTAINS THE HARDNESS DATA, CHEM1.DAT CONTAINS THE CHEMICAL DISSOLUTION DATA AND EXFreeEn.dat WHICH CONTAINS THE EXCESS FREE ENERGY OF SOLUTION OF INDIVIDUAL ELEMENETS IN ALPHA IRON. THE PROGRAM CAN WORK FOR AT THE MOST TERTIARY COATINGS. IN ORDER TO UPGRADE FOR MORE COMPLEX COATINGS ON HAS TO JUST ADD STATEMENTS RELATED TO THE ELEMENTS IN THE 'elea‘ ARRAY. ALSO, FOR NON STOICHIOMETRIC COATINGS ONE HAS TO ENTER THE NUMBERS WITH AN INTEGRAL NUMBER OF ATOMS IN THE MOLECULE. The following unit reads the file 'COMP1.DAT' for the compound name and its elements. INPUT COATING AND ITS DETAILS write(*,30) format('What is the coating?') read(*,3l) coating format(A8) open(unit=2,file='COMP1.DAT',blank='ZERO',status='OLD') read(2,53,end=68) elea(l),elea(2),elea(3),xa(l),xa(2),xa(3),compound format(A2,A2,A2,12,12,12,A8) if(compound.eq.coating) then goto 70 else endif goto 52 write(*,69) format('SORRY COATING IS NOT TO BE FOUND IN THE LIST') stop close(2) call mechwear(coating) call chemwear(coating,elea,xa) end Listing of 'subroutine mechwear' THE FOLLOWING SUBROUTINE GENERATES THE MECHANICAL WEAR DATA FOR THE COATING BASED ON THE HARDNESS OF THE ABRASIVE AND THE COATING. BOTH TWO BODY AND THREE BODY WEAR DATA ARE GENERATED. THE DECLARATIONS ARE AS FOLLOWS: -Ctng STANDS FOR THE COATING. ALL VARIABLES WITH 'COMP' ENDING DENOTE COMPOUNDS, ANY VARIABLE WITH h IN IT DENOTES HARDNESS.THE HARDNESS OF ANY MATERIAL CONCERNED IS EXPRESSED IN THE FORM H(T)=HO*EXP(-ALP*T). THIS WOULD THEN REPRESENT THE SOFTENING IN THE MATERIAL AT A HIGHER TEMPERATURE. ALL 59 60 C REPRESENT TEMPERATURE. THE SUFFIX 'ref'STANDS FOR THE REFERENCE COMPOUND AND THE SUFFIX C 'ab' STANDS FOR THE ABRASIVE. TO CHANGE THE ABRASIVE FOR TEH MODEL ONE HAS TO C CHANGE THE FIRST ELEMENT IN THE MECH1.DAT FILE. SINCE THE HARDNESS OF A MATERIAL C MAY NOT ALWAYS BE EXPRESSIBLE IN THAT FORM, AN INTERMEDIATE TEMPERATURE C IS CHOSEN SO THAT FROM THAT TEMPERATURE ONWARDS, THE HARDNESS TAKES A C DIFFERENT FORM. HENCE THE TWO TEMPERATURES t1 AND t2. subroutine mechwear(ctng) character comp*8,refcomp*8,ctng*8,abcomp*8 real h1,alp1,h2,alp2,refhl,refalp1,refh2,refalp2 real ahl,aa1pl,ah2,aalp2 integer t1,t2,reftl,reft2,t,i,at1,at2,abrat real th,ah,refh,relwear,abrwearcoeff,twobodyrelwear,abrwearcoefftwobody C IN THE MECHl.DAT FILE, THE FIRST RECORD CORRESPONDS TO THE REFERENCE TOOL MATERIAL C FOR THE RELATIVE WEAR CALCULATION. THE SECOND CORRESPONDS TO THE ABRASIVE PARTICLE. C HENCE THE ORDER OF DATA COLLECTION FROM THE FILE. c GETTING THE DATA FOR THE REFERENCE TOOL MATERIAL open(unit=2,file='MECH1.DAT',blank='NULL',status='OLD') read(2,500) refcomp 500 format(A8) read(2,510) refh1,refalp1,reft1 510 format(F7.1,FlO.8,I4) read(2,520) refh2,refalp2,reft2 520 format(F7.1,F10.8,I4) C GETTING THE DATA FOR THE ABRASIVE MATERIAL read(2,521) abcomp 521 format(A8) read(2,522) ah1,aalp1,at1 522 format(F7.l,F10.8,I4) read(2,523) ah2,aalp2,at2 523 format(F7.1,F10.8,I4) close(2) C THE NEXT UNIT RELATES TO SERACHING THE DATA FOR THE CANDIDATE TOOL MATERIAL. AS MENTIONED C BEFORE, HARDNESS IS EXPRESSED IN THE SAME FORM AS BEFORE. AGAIN FOR MORE CORRESPONDENCE C WITH THE EXPERIMENTAL VALUES, THE HARDNESS FORMULA IS SPLIT INTO TWO REGIONS. c LOCATING THE HARDNESS DATA FOR THE T/M IN THE FILE open(unit=2,file='MECH1.DAT',blank='NULL’,status='OLD') 525 read(2,530,end=560) comp 530 format(A8) readi2,540) hl,alpl,tl 540 format(F7.1,FlO.8,I4) read(2,550) h2,alp2,t2 550 formatiF7.1,F10.8,I4) if(comp.eq.ctng) then goto 580 else endif goto 525 S60 write(*,570) 570 format('SORRY ABRASION DATA NOT FOUND FOR THE TOOL MATERIAL') return 580 close(2) C THIS IS THE MAIN UNIT WHICH GENERATES THE WEAR DATA. AN OUTPUT FILE IS OPENED FOR WRITING C AND HEADINGS ARE WRITTEN. A SEPARATE SUBROUTINE 'hardness' FOR THE HARDNESS CALUCULATION USING THE C FORMULA IS WRITTEN. THAT WILL BE SEPARATELY EXPLAINED. APPROPRIATE ERROR MESSAGES ARE C ALSO WRITTEN SIDE BY SIDE. 61 openlunit=3,file='mechwear.out',blank='NULL',status='NEW') writel3,593) 593 format('Temperature(C) Relative Abrasive wear Two body relative wear T/Hardness A/Hardness wear coeff(3-body) Wear coeff.(2-body).') do 800 i=l,l4 t=(i-l)'lOO if(t.le.reftl) then refh=hardness(refhl,refalpl,t) elseif(t.le.reft2) then refh=hardness(refh2,refalp2,t) else write(*,600) 600 format('SORRY TEMPERATURE EXCEEDED FOR THE REF. MATL.:Termination') return endif if(t.le.t1) then th=hardness(hl,alp1,t) elseif(t.le.t2) then th=hardness(h2,alp2,t) else write(*,610) 610 format('SORRY TEMPERATURE EXCEEDED FOR THE T/M: Termination') return endif C THIS UNIT MEASURES THE ABRSIVE TEMPERATURE. AS SHOWN IN THE NEXT LINE A MODEST ASSUMPTION C THAT THE ABRASIVE TEMPERATURE IS 90% OF THE INTERFACE TEMPERATURE IS MADE. THIS IS AN C IMPORTANT POINT AND IS QUESTIONABLE. abrat=90*(i-l) if(abrat.le.atl) then ah=hardness(ahl,aalp1,abrat) elseif(abrat.le.at2) then ah=hardness(ah2,aalp2,abrat) else write(*,630) 630 format('SORRY,TEMPERATURE EXCEEDED FOR ABRASIVE MATERIALzTermination') return endif C A SEPARATE SUBROUTINE TO CALCULATE THE RABINOWICZ WEAR RATE CALCULATION IS WRITTEN. ITS NAME IS C 'volume'. THE WORKING OF THE SUBROUTINE IS EXPLAINED IN DETAIL LATER ON. 'abrwearcoeff' IS THE C VARIABLE WHICH CORRESPONDS TO THE 3 BODY WEAR. THE TWO BODY WEAR HAS A VERY SIMPLE FORMULA AND C HENCE NO SUBROUTINE IS NEEDED FOR THAT. abrwearcoeff=volume((th/ah),th) relwear = volume((th/ah),th)/volume((refh/ah),refh) twobodyrelwear = refh/th abrwearcoefftwobody=l/th write(3,640) t,relwear,twobodyrelwear,th,ah,abrwearcoeff,abrwearcoefftwobody 640 format(I4,' ',E9.4,' ',E9.4,' ',F7.1,' ',F7.1,' ',E9.4,' ',E9.4) 800 continue close(3) return end A.l.2 Listing of 'subroutine hardness' C THIS SIMPLE FUNCTION DETERMINES THE HARDNESS OF THE MATERIAL WITH THE ASSUMED FORM A.l.3 C ARE C BETTER. A.l.4 C INPUTS C THE 62 H(T)=HO*EXP(-ALPHA*T). THE INPUTS FOR THE SUBROUTINE ARE H0, ALPHA AND T real function hardness(h,alp,temp) real h,alp,rt integer temp rt=real(temp) hardness=h*exp(-alp*rt) return end Listing of 'subroutine volume' THIS SIMPLE FUNCTION DETERMINES THE RABINOWICZ WEAR RATE CALCULATION. ITS INPUTS THE HARDNESS RATIO AND THE TOOL HARDNESS. THIS IS ELABORATED IN THE LITERATURE real function volume(ratio,hard) real ratio,hard ifiratio.lt.0.8) then volume = 1/(3*hard) return elseif(ratio.lt.l.25) then volume = (exp(-2.S*alog(ratio)))/(5.3*hard) return elseif(ratio.gt.1.25) then volume = (exp(—6.0*alog(ratio)))/(2.43*hard) return else endif return end Listing of 'subroutine chemwear' THIS IS THE MAIN SUBROUTINE WHICH COMPUTES THE CHEMICAL DISSOLUTION WEAR. ITS ARE THE COATING, ELEMENTS ARRAY AND THE STOICHIOMETRIC NUMBER. subroutine chemwear(ctng,ele,x) character refcomp*8,ctng*8,comp*8,ele(6)*2,elem*2 real refmlv,mlv,refll,refm1,refnl,ref12,refm2,refn2 real ll,ml,n1,12,m2,n2 integer reftl,reft2,tl,t2,freen integer x(6),refx(6) integer free(6),refree(6) integer i,t real sol,refsol,relsol THE REFERENCE COMPOUND FOR THE RELATIVE WEAR CALCULATION IS THE FIRST ELEMENT IN CHEMl.DAT FILE. HENCE THIS CAN BE ALTERED SUITABLY. READING THE DATA FOR THE REFERENCE COMPOUND open(unit=2,file='CHEMl.DAT',BLANK='ZERO',status='OLD') read(2,100) refcomp,refmlv format(A8,F7.2) read(2,110) refll,refml,refnl,reft1 format(FlO.1,F7.2,F7.2,I5) read(2,120) ref12,refm2,refn2,reft2 format(FlO.1,F7.2,F7.2,IS) close(2) THIS UNIT COLLECTS THE DATA FOR THE CANDIDATE COATING FROM THE FILE. LOCATING DATA FOR THE INPUT COATING FROM THE DATA FILE 63 open(unit=2,file='CHEM1.DAT',blank='ZERO',status='OLD') 160 read(2,170,end=198) comp,mlv 170 format(A8,F7.2) read(2,180) ll,ml,n1,t1 180 format(FlO.1,F7.2,F7.2,IS) read(2,190,iostat=ios) 12,m2,n2,t2 190 format(FlO.l,F7.2,F7.2,IS) if(comp.eq.ctng)then goto 200 else endif goto 160 198 write(*,199) 199 format('SORRY DATA NOT AVAILABLE. COMPUTATION HALTED') return 200 close(2) C IF THE MOLAR VOLUME IS ZERO, THEN THE COMPUTATION IS HALTED FOR OBVIOUS REASONS. C TO SEE IF THE MOLAR VOLUME VALUE IS ZERO OR NOT if (m1v.eq.0) then write(*,201) 201 format('SORRY, THE MOLAR VOLUME IS ZERO. COMPUTATION HALTED') return else endif C THE NEXT SET OF THREE UNITS LOCATE THE EXCESS FREE ENERGY OF SOLUTION OF THE ELEMENTS C IN THE COATING. AT PRESENT THREE ELEMENTS CAN BE PRESENT. IF ONLY TWO ELEMENTS ARE PRESENT C THEN THE THIRD IS DUMMIED 'X' AND THE DATA CORRESPONDING TO IT IN THE FILE IS A ZERO. C TO LOOK FOR THE EXCESS FREE ENERGY OF SOLUTION IN ALPHA IRON FOR THE ELEMENTS open(unit=2,file='ExFreeEn.DAT',blank='ZERO',status='OLD') 205 read(2,210) elem,freen 210 format(A2,I7) if(elem.eq.ele(l)) then free(l)=freen goto 215 else endif goto 205 215 close(2) open(unit=2,file='ExFreeEn.DAT',blank='ZERO',status='OLD') 206 read(2,211) elem,freen 211 format