3 550 I HS. I LIBRARIES ’W . , MICHIGAN STATE UNIVERSITY J V EAST LANSING, MICH 48824-1048 , '3 f/ “CW5 This is to certify that the dissertation entitled DISSOLUTION WEAR: DECOMPOSITION OF TOOL MATERIAL, AND CONCENTRATION PROFILE INTO CHIP presented by Tim Kong-Ping Wong has been accepted towards fulfillment of the requirements for the Ph.D. degree in Materials Science and Mechanics 5 Major Pro;_essor’s Signature mks/ca Date MSU is an Affirmative Action/Equal Opportunity Institution PLACE lN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 c:/ClRC/DateDuo.p65—p. 1 5 DISSOLUTION WEAR: DECOMPOSITION OF TOOL MATERIAL, AND CONCENTRATION PROFILE INTO CHIP By Tim Kong-Ping Wong A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Materials Science and Mechanics 2004 ABSTRACT DISSOLUTION WEAR: DECOMPOSITION OF TOOL MATERIAL, AND CONCENTRATION PROFILE INTO CHIP By Tim Kong-Ping Wong The predictive capability of Kramer’s theory of dissolution wear has raised questions as to whether dissolution or diffusion is the dominant mechanism underlying tool wear. The diffusion mechanism finds support in experimentation and modeling of mass transfer (e. g., the Molinari-Nouari model of the experimental findings of Subramanian et al.), and fits more naturally in the setting of the irreversibilities of turning processes. Kramer’s dissolution mechanism, by contrast, is based on equilibrium thermodynamics. Our present aim is to re-phrase the dissolution hypothesis of tool wear as a boundary condition for species transfer within the chip’s bulk via diffusion. In this setting, dissolution is defined as the combined events of tool decomposition at the interface and the subsequent mass transfer of decomposed elements into the chip region. A set of equations is constructed wherein the tool is treated as a mechanically rigid, but chemically active, thermal conductor; the chip is treated as a rigid-perfectly plastic, thermally conducting material; and the behavior of the frictional-contact interface between tool and chip is described by a weak coupling of thermal and mass transfer. Chemical equilibrium is invoked for the distribution of tool species at the tool-chip interface. The Frank-Tumbull mechanism of molecular reaction (between interstitial impurities and vacancies to form substitutional impurities) is used as a hypothesis to explain the concentration profile of tool constituents into the chip as found by Subramanian et al. in 1993. The present interpretation of the Frank-Tumbull mechanism is illustrated by finite-element simulations. Physical parameters have either been estimated or curve-fitted to observed data, and the level of inexactness inherent in the curve-fit process limits the present work to a demonstration of only the sufficiency of the model. The assumptions made, their implications, parameters varied, and quantities calculated are summarized in the conclusion. ACKNOWLEDGh/fliNTS I am deeply thankful for the factors that have collectively made this moment possible. My thanks go to my family and friends in California, who have been the foundation of everything I do; to the admissions committee at Michigan State, for giving me the opportunity to pursue my goal; to the professors from whom I have taken classes, for their honesty and energy; to the administrative staffs at the Department of Materials Science and Mechanics and the Department of Mechanical Engineering, for their consistent help that at times reach beyond what is required of them; to my friends in Michigan, for all of the intangibles; and to the professors on my dissertation committee, for their grace and generosity in giving constructive criticisms. A semester of teaching assistantship that the Department of Materials Science and Mechanics has provided is gratefully acknowledged. The summer internship at NIST has been a treasured experience, not only from the point of view of learning and exploration, but also from the friendships that I have made. I have also benefited from the support of a unique organization—the Manufacturing Research Consortium at Michigan State University. My appreciation for what it has set out to accomplish is heart-felt, and I thank the organizers of the Consortium, as well as the sponsors who have visited our campus. Most of all, I would like to thank Dr. Patrick Kwon, for so many things. iv TABLE OF CONTENTS LIST OF TABLES .................................................................................. vii LIST OF FIGURES ............................................................................... viii KEY TO VARIABLES, PARAMETERS, AND OPERATORS .............................. x INTRODUCTION ................................................................................... 1 CHAPTER 1 THERMOMECHANICS OF TOOL AND CHIP WITH MASS TRANSFER ............. 6 The fully coupled equations ............................................................... 6 Weak coupling of the equations of plasticity and mass transfer ....................... 10 CHAPTER 2 THE ZEROTH—ORDER PROBLEM FOR THE TEMPERATURE BACKGROUND ..12 Aim of 2D simulations .................................................................... 14 Model and governing equations ........................................................... 14 Finite element discretization ............................................................. l7 Implementing the interface conditions ................................................... 20 Uncoupled interface ................................................................... 22 Temperature continuity ................................................................ 22 Convective cooling (and flux continuity) .......................................... 23 Mixed interfacial behavior .............................................................. 24 Simulation results ........................................................................... 24 Mixed interfacial behavior ............................................................ 26 Temperature fields ........................................................................ 28 Summary of the temperature problem ................................................... 29 CHAPTER 3 THE FIRST-ORDER PROBLEM OF SPECIES TRANSFER ......................... 30 The time-dependent equations of the Frank-Tumbull reaction and mass transfer .. .30 Boundary conditions for mass transfer ............................................ 32 Connection to previous work by Kramer .............................................. 33 The time-independent equations .......................................................... 34 The difference between W and Co profiles; non-dissolution mechanisms ......... 34 Results of the species-transfer problem .................................................. 37 Unusually low values of mass diffusivities ......................................... 41 The determination of interstitial solubility; post-machining relaxation ........ 44 Solubility as the maximum amount of dissolution wear achievable ............ 48 Amount of worn cobalt not due to dissolution wear ................................. 49 The rate of tool wear as a field quantity ............................................ 53 Table of Contents (continued) Measurement methods for checking the distribution of vacancies ..................... 55 N on—uniqueness of Frank-Tumbull representation .................................... 56 Notes on the numerical method .......................................................... 57 CHAPTER 4 CONCLUSIONS AND SUMMARY .......................................................... 59 A list of assumptions made, with comments ............................................ 59 A list of key parameter variations used to achieve match with experimental data . . .65 A summary of motivations, results, and quantities calculated ...................... 66 APPENDDC .................................................................................... 68 BIBLIOGRAPHY ............................................................................... 77 vi LIST OF TABLES Table 1: Fully coupled equations ............................................................ 8 Table 2: Weakly coupled equations .......................................................... 9 Table 3: Parameters used in the Oth-order problem (background temperature) ....... 26 Table 4: Parameters used in the lst-order problem (mass transfer) .................... 41 vii LIST OF FIGURES Figure 1: Concentration profiles of tool constituents into the chip, reproduced from Subramanian et a]. [1993]; chips are quenched after machining to preserve profile shape. The schematic illustrates the cutting geometry. .............................................. 2 Figure 2: Example of humped distribution with arbitrarily chosen functions ......... 3 Figure 3: Sources for and location of the Frank-Tumbull reaction ....................... 4 Figure 4: Decomposition of fully coupled equations ....................................... 7 Figure 5: Schematic of cutting geometry; boundary conditions of heat transfer ........ 14 Figure 6: Ideal l-D representation of the work and tool materials ....................... 20 Figure 7: Mesh used for O-th order problem, and a magnified view ..................... 25 Figures 8a, b: Comparisons of heat partitions of mixed interfacial conditions to the case of pure temperature continuity ................................................................ 27 Figures 9a, b: Interface temperatures corresponding to Figures 8 ........................ 27 Figures 10a, b: Temperature fields (see also Figures 9a, b) ................................. 28 Figure 11: Schematic of the Frank-Tumbull reaction ....................................... 30 Figures 12a, b: Experimental concentration profiles, deduced from Subramanian et al [1993]. Empty and filled circles denote data points. In the original data, the 240-m/min curve in (b) asymptotes to the same non-zero value as the lSO-m/min curve. .......... 35 Figure 13: Variation of temperatures away from the average ............................. 38 Figure 14: Mesh used in the lst-order, species-transfer problem ......................... 38 Figures 15a, b: Sample X V and X i distributions ......................................... 39 Figures 163, b: A comparison of available values of diffusivities versus actual values used in simlations for (a) 2.5-m/s cutting speed, (b) 4-m/s cutting speed .............. 42 Figures 17a, b: X V and X i distributions using table-lookup values of Figures 16 .....43 Figure 18: Good match in an early run that turns out wrong ............................ 45 viii List of Figures (continued) Figures 19a, b: Why the good match of Figure 18 does not work ........................ 45 Figures 20a, b: Attempts at matching experimental W-concentration profiles using only steady-state W distributions from the model ............................................. 46 Figures 21: (a) Relaxation of X s after cutting; (b) X s after 3.6 millisecs. of relaxation ..................................................................................... 47 Figures 223, b. A check on the exhaustion of tungsten X i during relaxation ........ 48 Figures 23a, b: A check of Kramer’s hypothesis that dissolution wear is bounded by solubility ......................................................................................... 49 Figures 24a, b: The relaxation process. A check on the exhaustion of cobalt X i during relaxation .......................................................................................... 50 Figures 25a, b: The W-Co ratio within the chip as residual X i ’s are exhausted during relaxation for (a) 4-m/s cutting speed, (b) 2.5-m/s cutting speed ........................ 51 Figures 26a, b: The surface rate constant for tool decomposition and tool-interface temperature .................................................................................. 54 Figure Al: Ideal l-D representation of the work and tool materials .................... 69 ix Key to Variables, Parameters, and Operators A, B : impurity and host material, respectively Ai & As , BV : interstitial & substitutional impurities, and host vacancies, respectively b : body force per unit mass (N/kg) bk : body force per unit mass for species k * c : total concentration of all species (mollm3) ck : molar concentration of species k * 6V : mass-based specific heat at fixed volume (J/(kg-K) Ework : workpiece EV (work & chip), Oth-order problem étool : tool EV , 0th-order problem C [j : ij-th component of the advection matrix in the Oth-order problem D : rate of deformation ( 1/s) Dk : diffusivity, within iron, of species It (mzls) * DOF: number of nodes in the Oth-order problem fk : rate of mass generation of species k per unit volume (kg/(m3-s)) * F k : (e.g., Fs , Fv, psz ) rate of mole generation of species k per unit volume (mol/(m3-s)) * F i]- : rj-th component of the effective flux- coupling matrix in the 0th-order problem Fem , chcd , Fshear , kac : cutting forces for the Oth-order problem (N) g1 , E1 : lst-order Gibbs free energy (see Notes) AEactivn : change in molar Gibbs free energy of chip—boundary activation (Jlmol) agdccomp : change in molar Gibbs free energy of tool decomposition G : Gibbs free energy (I) h : heat-transfer coefficient (W/(m2-°C)) hi : bilinear shape function over node i I : the union of the rake and flank interfaces jk : diffusive mass flux for species k t with respect to v J k * : diffusive molar flux of species k * with respect to v" k : thermal conductivity (W/(m'K)) kwork : workpiece thermal conductivity (work & chip) for the Oth-order problem ktool : tool thermal conductivity for the Oth- order problem kfwd : rate constant for the forward Frank- Turnbull reaction (l/s) kdiss : rate constant for interfacial dissolution kV : rate constant for vacancy production K ij : rj-th component of the thermal-conduction matrix, Oth-order problem (digs : length scale for interface dissolution (m) L, Lwork , Ltool : length of the work or tool element in the one-dimensional heat-transfer example Lmkc : tool-chip contact length along rake (m) m : total mass of all species (kg) mk : mass of species k * M k : molecular weight of species k (kg/mol) * M i]- : ij-th component of the heat-capacity matrix, 0th-order problem N : total number of species p0: 0th-order pressure (N/mz) q : heat flux (Jl(m2's)) qo , q] : Oth- and lst-order heat fluxes qflank , ‘Irake : heat fluxes along the flank and rake interfaces, OtIi-order problem Qi : i-th component of effective external forcing in the Oth-order problem (W/m) Q: interfacial heat flux in the one-dimensional heat-transfer example le , sz : end-node heat fluxes for the work, one-dimensional heat-transfer example Qt] , Qtz : end-node temperatures for the tool, one-dimensional heat-transfer example r : rate of heat generation per unit mass (Wlkg) ’work : rate of heat generation within the primary shear zone, 0th-order problem R : universal gas constant (cal/(mol-K)) R,- : i-th component of effective internal forcing in the Oth—order problem (W/m) t : time (5) t1, t2: end nodes of the tool in the one- dimensional heat-transfer example ti]- : ij-th component of tool ‘stiffness’ matrix, one-dimensional heat-transfer example T : Cauchy stress (N/mz) u, r? , r7 : internal energy (J/#) (see Notes) Key to Variables, Parameters, and Operators (continued) 130 , ii, , r71 : 0th- and lst-order internal energies (see Notes) v : mass-averaged velocity (m/s) v k : velocity of particles of species k t v*: concentration-averaged velocity v x , v y : horizontal and vertical workpiece velocities, 0th-order problem cht . ered . Vshear . Vrake : cutting speeds for the 0th-order problem V, I? (= p.1 ), I7: volume (see Notes) w : Galerkin weighting function w1, w2 : end nodes of the workpiece in the one- dimensional heat-transfer example wij : ij—th component of workpiece ‘stiffness’ matrix, one-dimensional heat-transfer example W: width of system transverse to flow (m) x : horizontal coordinate (global or local) Xk (i.e., Xi , X5 , Xv ): mole fraction of speciesk (=ck lc) * X k, cq : solubility, within iron, of species k t X Co, cq : solubility of cobalt in iron X W, eq : solubility of tungsten in iron y : vertical coordinate (global or local) Yk : mass fraction of species k (= mk /m )1: a: fraction of impurities within the chip lattice identified as interstitial impurities ,6: fraction of heat partitioned to the workpiece interface, Oth-order problem 17, I? : entropy (J/(K°#)) (see Notes) fio , fil : 0th- and lst-order entropies (see Notes) 9 6 9 a): fraction by which the tungsten solubility is multiplied to obtain the effective solubility of cobalt in iron during dissolution Vpc : Debye frequency of iron lattice p: average density (kg/m3) pk : density of species k * Pwork : workpiece (work & chip) density, 0th- order problem Ptool : tool density, 0th-order problem 9: temperature (K or °C) 00 , 01: 0th- and lst-order temperatures 6 j : j-th node temperature, Orb-order problem Hwork : workpiece (work & chip) temperature, 0th-order problem atool : tool temperature, Oth-order problem 6w 1 , 9w2 : end-node temperatures for the work, one-dimensional heat-transfer example 6“ , 60 : end-node temperatures for the tool, one-dimensional heat-transfer example w, r]? : Helmholtz free energy (see Notes) #70 . V713 Oth- and lst-order Helmholtz energies A difference operator V spatial gradient V- spatial divergence V-V Laplacian 8(-)/8t partial derivative with respect to time D(-)/Dt total (or material) derivative with respect to time T1 :T2 = trace(TlTT2 ), tensor product (V :T)-c = V - (TTc), divergence of a tensor T, where vector c is any constant vector * Index k takes on the values of ‘i , s , or ‘V’, denoting ‘interstitials’, ‘substitutionals’, or ‘vacancies’, respectively, within the chip lattice in the lst-order (species transfer) problem. Notes: Extensive quantities such as volume V, internal energy u, entropy 7], etc., are written on either a mass or mole basis. For uniformity in notation, an overhead caret (e.g. r3 ) is used to denote a quantity per unit mass, whereas an overhead bar is used to denote a quantity per unit mole (e.g. 17 ). Following standard notation, partial molar quantities is denoted by a subscript indicating the particular species (e. g., 170, = Bu ldna ). The ‘#’ symbol denotes a wild-card unit that could either be kg or mol depending on whether a quantity is expressed on a per-unit-mass or per-unit-moles basis. Quantities that are treated in the numerical simulations as material constants are not italicized (e.g., thermal conductivity, k). INTRODUCTION Dissolution as a high-temperature wear mechanism has first been proposed by Kramer [1979]. The key idea is to interpret the solubility of a tool material within the work piece as a measure of the tendency for dissolution. A condition of chemical equilibrium involving free energies [Kramer, 1979; Kramer and Suh, 1980], namely, AGdecomposition of tool = AGdissolution of tool components into work is introduced to estimate values of solubility. In this context, ‘dissolution’ is taken as the placement of tool components into the work. Further, a simple argument relates the solubility to the volumetric wear rate up to an unknown (but assumed common) multiplicative factor, which enables the life expectancy of different tool materials to be compared on a relative basis. The quantitative results not only conform to common experience, but also identify certain nitrides and oxides for their dissolution-wear resistance at high-temperatures. Hence, in contrast to the theories at the time that favor the non-equilibrium mechanism of diffusion [Cook and Nayak, 1966; Naerheim and Trent, 1977], the importance of an equilibrium quantity, namely the solubility, is thrust to the fore. Over 20 years after the seminal work of Kramer, the actual concentration profiles of tool constituents dissolved into the chip have been measured and reported by Subramanian, Ingle and Kay [1993]. Their data, which correspond to the machining of an A181 1045 steel using a tungsten-carbide tool with cobalt binder (W C-Co), are reproduced in Figure l: 3 E Original data: mass fraction of W in chip 0 ‘ Original data: mass fraction of Coin chip A .J ' A .I4 g A 5 3 r f, 0.12 ‘5 240 rn/min ‘5 240 m/min a 2.5 ‘ O) 0 1 .8— 2 ._ o 08 E V . g 1‘5 150 m/min ‘ €006 150 m/min E 1 $0.04. I 0.5) '33, 0.02 i a - o 0.1 . 0.2 ‘ 0.3 0.4 0 0.1 52 0.3 0.4 depth into Chip (Hm) depth into chip (pm) Tool-chip interface , .. ,~ 7. , ; Fi ure 1. Concentration rofiles éééééééfig’é; c 'p ow of tool constituents (tungsten (W) // / 4,0 : jéflé/éé/ ; and cobalt (00)) into the chip, 4%; 2%; i . 3%: Tool 3%? . re reduced from Subramanian et 574' 'éé p ”/3 :/x , . igégzéfiéééé : concentration al. [1993]; chips are quenched éfié/ééy’ééflé . profile . . gggégggggg : after machining to preserve ¢¢¢//¢ /¢j// ; _ 232%?Zggézéé 5 profile shapes. The schematic ZW/Wfléw . M/zéiéééffi : illustrates the cutting geometry. <:Il work flow According to the authors, the humped distribution may be due to the formation of a dislocation channel close to the contact interface. As seen from Figure 1, such a dislocation channel would have a width of about 1/10 of a micron and is located about 1/10 of a micron away from the tool-chip interface; this is compared with a total depth of a chip that measures hundreds of microns. As the regions adjacent to the primary and secondary shear zones also experience severe shear deformation, it is not easy to see why the dislocation channel is so precisely narrow in its appointed location. Solubility alone cannot explain the humped distribution either, for a humped solubility profile would require a humped temperature profile away from the interface, which stands in contrast with an exponentially decaying behavior that is well known (see also Chapter 2). Using a set of coupled, reaction-diffusion equations, Giese, Stolwijk and Bracht [2000] have shown that a humped profile of copper and nickel impurities diffusing in a germanium lattice can be explained by the Frank-Tumbull mechanism of molecular mass transfer [Frank and Tumbull, 1956; Sturge, 1959]. Delaying the discussion of this mechanism until later, it suffices to say that the enabling idea is to derive a humped distribution as the product of a decaying function and an increasing function, as the following example illustrates: 20 __ 1 .5-30x+175x2 Figure 2. Example of humped distribution with arbitrarily chosen functions We use the same idea to capture the data of Subramanian et al. The reaction-diffusion equations of Giese et al. are supplemented with a simplified model of advection to describe mass transfer in the presence of chip flow. An exponentially decaying, interstitial profile is obtained by solving an advection-diffusion equation, while an increasing vacancy distribution is chosen so that its product with the interstitial decay— under the requirement of mass conservation of the Frank-Tumbull reaction—fits the experimental data. The result is an interpretation of the humped profile via the Frank- Tumbull mechanism. The role of chip flow is to establish thermal equilibrium, and to transport the vacancies created in the primary shear zone directly into the chip, where they interact with the tool interstitials in the proposed Frank-Tumbull reaction; this is summarized in Figure 3. Chip U tool-chip interface: location of source of interstitials F rank-Tumbull / reaction tool primary shear / zone: creation of work Q vacancies Figure 3. Sources for and location of the Frank-Tumbull reaction Kramer’s original idea on solubility [Kramer, 1979] is re-interpreted within the context of the rate constant of dissolution, which in turn enters the boundary conditions of the proposed advection-diffusion-reaction equations. It is postulated that the boundary flux of dissolved components across the tool-chip interface is proportional to the available solubility; the mole fraction of decomposed tool constituents; and the mole fraction of vibration-activated openings on the chip surface that are wide enough to admit tool constituents (also called “activation complexes”). A similar rate expression has been discussed by Suh [1986, p. 369], which is less restricted than the present one in that in Suh’s expression, the idea of activation complexes is not needed. In the present interpretation, dissolution is defined as the combined events of the decomposition of tool constituents at the interface, and the absorption of these constituents by the chip. That diffusive mass transfer does not account for the entirety of tool wear can also be motivated from the profile data of Subramanian et al.; this idea is introduced in Section 3.4 and re-visited in light of the present results in Section 3.5.4. The equations of mass transfer are incorporated into the balance laws of mechanics in a simplified way. First, the balances of mass and linear momentum are satisfied using a chosen advection field and a simple force balance (i.e., classical machining theory). Second, the balance of energy is decomposed additively into a zeroth-order problem governing the background temperature, and a first-order problem governing the mass transfer from the tool into the chip. The additive decomposition of energy is based on the large difference in time needed to achieve a steady state for thermal diffusion versus mass diffusion (Section 1.2). Bridgman’s notion of generalized entropy [Bridgman, 1950] gives an operational meaning of temperature that appears in the energy equation (Section 1.1). The general organization of the dissertation is as follows. Chapter 1 contains a review of the coupled governing equations and an explanation of the decomposition into the temperature problem and the species-transfer problem. In Chapter 2, the temperature background problem is solved. Chapter 3 contains the main results of the dissolution- wear problem as stated before, including a discussion of the non-uniqueness of the Frank- Turnbull representation, and a brief survey of experimental methods that are available for measuring the spatial distribution of vacancies. Lastly, Chapter 4 is a summary of results. Chapter 1 THERMOMECHANICS OF TOOL AND CHIP WITH MASS TRANSFER A model of particle exchange from one region to another in the presence of bulk motion and deformation involves a sizable number of coupled equations. However, for the high—temperature phenomenon of dissolution considered here, we place focus on the solution for the temperature field and the tracking of species exchange, while those equations associated with the purely mechanical aspects are only approximately satisfied. The result will be a set of equations involving advection, diffusion, and reaction (Chapters 2 and 3). 1.1. The fully coupled equations The tool is treated as a mechanically rigid, but chemically active, thermal conductor, and the only equation of concern is the energy equation, or Equation (p) of Table 1 written for the tool variables. For the tool, the stress power T : D = 0 because of rigidity. The chip is treated as a rigid-perfectly plastic, thermally conducting material, and all equations of Table 1 are applicable. In the present approach, the thermodynamics of the cutting process is not predicated on the accessibility of states via reversible processes, from which entropy is constructed. Instead, by appealing to the recoverability—via a hysteresis process—of any point in a perfectly plastic process, a generalized entropy is postulated to be a state function [Bridgman, 1950]. Further, by arguing that the material’s structure is unchanged by planes of perfect slip, a generalized, system entropy is chosen to be the same one that is used for a material that is rigid at all locations except planes of perfect slip. Irreversibility is indicated by a 100% conversion of the plastic work into heat; in other words, T : D is treated as a heat-source term, and strain does not enter explicitly as an argument of the energy functions. The escaped heat from plastic work adds to the entropy of the surrounding material, and is the basis through which an absolute temperature of the system is prescribed in the presence of perfectly plastic deformation. Such a temperature is precisely the same one that enters the energy equation (60f Equation (p), Table 1). main variables: main variables: “x, chip R90» ‘70- P0 Xi. Xv. Xs ’IO 0 o -:> _ I: g g Q + ’1; o c .3 work o g é I O o A 90, V0, p0 are , order problem Oth-order problem: let-order problem: coupled heat-transfer only species-transfer only (Table 1) (Table 2) (Table 2) Figure 4. Decomposition of fully coupled equations Table 1 summarizes the governing equations of the fully coupled problem. For discussion and derivation of the equations, the reader is referred to the references [Mase and Mase, 1999, Ch. 5; Bird et al., 2002, Ch. 19]. In spite of the numerous equations, Figure 4 shows the simplification that an additive decomposition brings and the key variables that are of interest. This decomposition is justified in the next section. Ta 0‘ Ab «HIM! + m D 0T u a “Lommcab “so: :8 32 3:25.80 I 2 AS OVQII QIAU aDrngnQE—WIIQQIIIQIQ "339503 :Q «Q SQ I I; Q: Ad. I: .DH. «Isl N +9. a. + a. .D HQIIQ. wQ .33 ame_m\zw:0:m 5 SW" Q I v. I u Q ..3 A8 «IMIQ SHTMQQ Q. 5+ D MIQQ chem A5 A (“6 I «a v . E u 430:0 00¢ 5—28—03 AA—sz....o—\fl 9%V%.E "A z» 3... —\~ afi V%.E "A ZXET... ~kEo§V$HAZEo:;—Eowvsn 5 . mm .5 I n : 239:5 o 00:03:: A A—lzxa...ouxmawvm}m ( M m a Q A: c n 0:— E03533. C: «a. saw +Q ” P+ a. DI :Q n ..I.QIQ .xw:0:m 3 @602 mfihfi 20 203 3 Mb 0 u. H h 608:3 E0520 58:35 3 .9 Q n aQ + a. u D "00:23 8:52:02 fl VAKHSAH.D+| .VMIQIQQ A5 0 u d. N 2935800 30:: mo 3:033:00 A3 A :m “AGREES 860mm S A.» I «i «Q n 0;. ”880% 5.« :8 56 03255 A8 «.3 «Q ”0.080% :3 :8 08c :38. a 0A A3 > Q "A a.» «3 N “.3622, nowfiofiémflz 3 Q n «QN “56:0: 0w0:0>< A V :0 28 2 Q :o . :Q >. s _ _ 585800: m cm s : m a a u . or QIQI .no .D Eb. Q C on A QQD+Qm Ea: oom~E=80>O 88.35 3950 :3: ._ 032. 3 A3 8 A3 A5 8 3 3 5 A3 A8 A00 :Q :Q 0 w AoQ.00.._Iz.:....._.c_mQ "0 5 0:0 5 ll QIE. I 0 Q 280:8 0.3 :09“: > "AT 2.... ...ESQ 0 m 5 :0. DI u A:I2.fi...._.c€Q Q ..Aw:0:m "E 5 0 29:80:00 00625 ... :I0u0.+ :Mx .0...» :8 A... > I a; «0 n .. 0:. ”008000 5-0— .:0 :38 508 030885 ...> 0 n A 0.» «0v N 09:020.: 00w0:0>0-:000:800:00 a a 0190..- z:.....5§00 ”330880 7...: Q AVQIAéDa. IQII” M+5DII _:QQQ ..0QI 9304658385." 3 006 I .I. A 00” 72.: 5 :8. ”3:05 008 20:8.0m & _n« «km 82—" AS «nfllwu ..Iflmll n :0 3:6 000$ l 2 2 8 USD «QI II. .. 0: _. £8808 0008 :8 30. 03.80800 AEV 0; «0 ”00:000.: 5-0. :0 56 8.02 C: 0 n «0N 8000:8088 00:0,“. 80308 :00:0-:00:0~ .3 0008 0:0 0Q 080008 0:0 . :IQ u 0> . 00 80 0030> 5 m 5 Q8808 0.00? :Q lQH30 :000 00 . 20...... i. m owe“: I 0.0.: n A 0$ 0% 23:05 008 BEE—om :Q H O . l"l ” Q P+ a D AonomQQ 3:05— ..330 0: ~00: 80%:0583: 005$ EwEEQ $034,058 :Q ” +2.. In >0Q ”3 :8: 3:0: :0 Q a. D 00 Q .. 3 h: m 0 . 2. .0 o q :8 c AV Q..—.+ .DIIAIPQmfimQ . :m A8 <0>.:... n >09" 8... 2:00:08 8080 .3 @0003 A8 ameI u 3: 338 :003 5:005 882008 30:80.0 0:.» £00: 3808 :0 :0 0080000 ”00:30: 8080882 0088008 0:0 880: 8:0 05 8 0:0 580: 0:03 05 8 8:88: 00—08 30: .3003 ”5:553:08008 0:0 b88800 0008 =0:0>O ”$338580: 839.8303 8030i :00:0-:Q:0:0N 800005 008000 A3003 .N 030:. 1.2. Weak coupling of the equations of plasticity and mass transfer The reformulation of the fully coupled equations into two simpler problems is given in Table 2. These two problems correspond to decomposing the variables and functions {6,fi,fi,t/?,q} additively as follows: {flail/1‘1}: {Hoifio’fio’Wofloh{91,771,fi1,l/71’¢11}’ (1) where the subscript ‘0’ refers to the zeroth-order problem of determining the background temperature, and the subscript ‘1’ refers to the first-order problem governing molecular mass transfer. The motivation for this additive decomposition is that first, the rise in system temperature is almost entirely due to plastic deformation and friction, or equivalently, molecular mass transfer from the tool does not significantly affect the background temperature.l Second, for a given distance, the steady state for thermal diffusion is established much faster than the steady state for mass diffusion, which gives the decomposition (Equation (1)) a natural division in time scale. In fact, the thermal diffusivity of a material is typically orders of magnitude larger than the mass diffusivities of impurities moving through the material. For example, at 900°C and atmospheric pressure, the thermal diffusivity of iron is on the order of 10'6 mzls, whereas the mass diffusivities of cobalt and tungsten in iron is on the order of 10’14 mzls [Elliott et al., 11963, p. 690 for Co, p. 695 for W]. For the zeroth-order problem, the balance of mass and the balance of linear momentum (Equations (a) and (i) of Table l) are satisfied by assigning a simple advection field and by balancing the cutting forces at steady state, as it is done in the 1 The system temperature is significantly increased by drastic changes in the tool geometry due to wear, but this temperature increase is not due to mass-transfer coupling, but rather to changes in the cutting forces because of a change in tool geometry. 10 classical machining theory [Shaw, 1984]. The only use of plasticity is through an idealized, 100% conversion of plastic work and friction into heat. The rate of plastic work is connected to the cutting forces by the work-energy theorem, Equation (e) of Table 2. The conversion of cutting forces into friction power is discussed further in Chapter 2. The first-order problem pertains to the small amount of mass transfer due to too] wear that is postulated to have no effect on the zeroth-order temperature and density fields. Hence, the quantities of interest are mass-transfer variations with respect to fixed background temperature 6]) and specific volume )7 = 170 , as indicated in Table 2 (Statement 0)). Equilibrium distributions are governed by the Second Law, which is also expressed in terms of the Gibbs free energy g, = 1/71— p090 (Inequality (v) of Table 2). The interfacial pressure p0, which can be approximated from dividing the cutting forces by the contact area (the present work will not go farther than requiring an order-of- magnitude approximation), are those of the zeroth-order problem. Additional relations in Table 2 (Equations (k) thru (q)) not found in Table l are discussed in [Bird et al., 2002, Chapter 19]. ll Chapter 2. THE ZEROTH-ORDER PROBLEM FOR THE TEMPERATURE BACKGROUND Because species transfer is so closely confined to a submicron region adjacent to the tool-chip interface, emphasis is placed on the details of the interfacial temperature distribution. However, the present calculations will also return the temperature field in a much wider region of the tool, work, and chip. In turning processes, while a large portion of the contact interface between tool and work is under intimate contact [Zorev, 1958 (Fig. 54.7 on Plate 5); Wallace and Boothroyd, 1964 (Fig. 14, p. 84); Trent, 1967] (and hence temperature continuity holds), usually a distinct downstream region of gross, material deposition can be found [Doyle et al., 1979; Ackroyd et al., 2001; Kim and Kwon, 2001]. It is in this latter region of looser contact that a temperature jump is conjectured, even at steady state conditions. The aim is to find plausible conditions for a cooler work interface in orthogonal turning, as it pertains not only to chemical wear but to abrasive tool wear2 as well. From the solution of the heat equation, it is known that an interfacial temperature jump results if a convective-cooling condition is used with a heat transfer coefficient. Furthermore, the adiabatic condition and temperature continuity are two extremes of 2 It is known that the hardness of metals and their inclusions decreases exponentially with increasing temperatures [Kramer and Judd, 1985]. In turn, abrasive wear has been shown to depend on the hardness ratio between tool and inclusions in the work material [Rabinowicz, 1977]. Abrasive wear intensifies if, during frictional contact, the temperature drops in going from tool to work material, as the inclusions in a cooler work material become effective abrasives. 12 behavior representing assignments of 0 and co to the heat transfer coefficient, respectively. Here, the convective medium is the moving workpiece; however, the situation is complicated by the presence of friction. In fact, the amount of frictional heat that is partitioned to either side of the interface becomes an independent unknown. This is studied in the Appendix through the use of a 1D problem, where it is found that the heat—partition function enters directly into the expression for the interface-temperature jump (Equation (A23)). Because of the lack of exact solutions for complex flux conditions and geometry in 2D, results derived in the 1D example serve only as guidelines in the 2D FEM simulations. It is noted that the interface conditions considered are linear in the boundary temperature, as opposed to a non-linear condition, such as radiation, for which the present results do not apply. While experimental proof of temperature jumps across frictional interfaces has yet been found, the results of the present study imply that if interfacial temperature jumps exist, then they can be characterized by the behavior of the heat-partition function across the interface. Assuming that temperature continuity is associated with intimate contact, the absence of intimate contact can then be quantified as deviations in the heat-partition function away from its value in the temperature-continuity case. For an extensive coverage and review of analytical work on calculating temperatures in machining, and in particular, on the relationship between temperature continuity and heat partitioning, the reader is referred to a set of papers by Komanduri and Hon [2000; 2001a; 2001b]. The present discussion is largely motivated by the work of Chao and Trigger [1955; 1958]; here, we are interested in the opposite problem of how heat partitioning can be manipulated to achieve temperature jumps. 13 The subscript ‘0’ denoting the zeroth-order problem is dropped for the remainder of this chapter. 2.1. Aim of 2D simulations It will be seen that 2D simulation results are consistent with exact 1D results found in the Appendix. Namely, for the part of the contact interface where the convective-cooling condition is Operative, if the heat partitioned to the work is less than that for the case of temperature continuity, then the work temperature will be cooler than the tool temperature (Inequalities (A24)). In turn, such a temperature jump will have ramifications for dissolution wear (Chapter 3). 2.2. Model and governing equations The two dimensional geometry of concern is shown in the following figure, which corresponds to orthogonal turning with a 5° rake angle: zero flux ambient chip j {V rake face tool { adiabatic “L7 IL adiabatic \ 13’ [ ambient 7 flank face \7: => f zero flux work Figure 5. Schematic of cutting geometry, boundary conditions of heat transfer The governing equations for the work and the tool temperatures are a re-phrasing of Equations (c) and (h) in Table 2 as follows: 14 2 2 [kwork[a 6work + a 6work] l a 39 3x2 2 Pworkcwork 301k :1 3y i (2) t A 36work aHwork " Pworkcwork Vx T + V y T + pworkr work . aa 320 826 Ptoolctool 5:01 = ktool[ 31:30] + aytgol] (3) Here, x is the horizontal direction in Figure 5, y is the vertical direction, r is the rate of heating in the primary shear zone, k is the thermal conductivity, p is the density, 6 is the specific heat (per unit mass), v is the cutting speed, and 6is the temperature. Equations (2) and (3) are coupled through conditions at the contact interface (marked by the letter I in Figure 5). Boundary conditions are as follows. Infinitely far away from the cutting edge and contact interfaces, the system temperature returns to the ambient. Downstream of the contact interfaces along the rake and flank faces, adiabatic boundaries are assumed; that is, the bulk of the heat transfer is due to conduction across the contact interface I, and radiation far away from I does not play a key role. For the moving work and chip, the outflow boundary condition for the advection-diffusion equation is that of zero flux (k - grad6= 0), as discussed in the text of Zienkiewicz and Taylor [2000b, pp. 44-46]. The advection term in Equation (2), .. 86 66 Pworkcwork[vx _£\;/Tork + Vy 7%] 9 (4) describes the flow of the work piece and its chip. For convenience, Equations (2) and (3) will be written in condensed form as p6%q-=kV-V6—pcv-V6+pr, (5) 15 where it is understood that if the equation refers to the tool, then the velocity v is equal to 0, and the internal source r equals 0, etc. The symbol V denotes the spatial gradient. Stress and strain fields are not direct inputs to the energy equation that is being considered. Instead, they enter through the rate of plastic work, which is considered as a heat-source term as stipulated in Chapter 1. In particular, the source term T : D in Equation (h) of Table 2 is related to the external power through Equation (e) in Table 2. The basic inputs are forces3 and velocities { Fcut , vcut } and { Ffeed, ered} in the cutting and feed directions, which can be resolved into { Fshear» vshwr } and { Fake, Vchip} along directions tangent to the primary shear zone and the rake face, respectively [Shaw, 1984, pp. 21-22]. The heat generated in the primary shear zone is expressed as a body source r work : Fshear"'shear ’ (6) while the frictional heat generated along the secondary shear zone (tool-chip, or rake interface) is expressed as a boundary flux, Frakchhip —— . (7) Lrakew (Irake = (q ' n)rake = where q and n are the heat flux vectors and outward normal to the boundary, respectively, Lrake is the length of contact, and Wis the out-of-plane thickness of the system. The frictional heat flux qflank generated along the tertiary shear zone (tool- work, or flank interface) is also expressed as a boundary flux. As implied by DeVries’ 3 Knowing the cutting geometry and the cutting speeds, forces can be converted from specific-energy data that is tabulated in DeVries’ text [1992, pp. 106-121] based on the classical machining theory. Here, we simply make use of the force and average-temperature data reported by Subramanian et a1 [1993]. 16 text [1992, Figure 5.3c, p. 113], the product of qflank and the flank-wear area can be obtained as the correction, in the presence of flank wear, to the total power input Fcutvcut + Ffeedvfeed’ (8) so that ‘Iflank = (q 'n)nank _ {change in chvcut + Ffeedeeed due to flank wear correction} . (9) Lflankw The quantity Frake in Equation (7) is calculated from Fcut and Ffeed before the flank- wear correction is applied. The flux sources qrake and qflank are then uniformly applied to the contact regions regardless of whether they are prescribed to be temperature- continuous or convective-cooling. The interface heat-transfer coefficient h is known to be highly dependent on process conditions. An intermediate value of h = 1000 W/(m2-°C) is used in the simulations.4 2.3. Finite-element discretization The finite-element weak form of the governing equations is derived by pre- multiplying Equation (5) (or Equations (2) and (3)) by a weight function w, integrating over the respective domains of work and tool, and performing integrations by part, resulting in 4 As a crude indication, a value of h = 400 W/(m2-°C) has been reported between molten aluminum and its die (http://msewww.engin.urrrich.edu/research/groups/pehlke/publications/lppm/index_html). 17 [w pég—fdA= -[(Vw).(kva)dA-[wpcv -vadA+[wprdA+[wkv0-nds (10) where I(-)dA denotes area integration, and is taken over the entire areas of tool and work, and n is the outward normal vector to the boundary element ds. Because of the simplifying assumption that the boundaries of the system are mostly adiabatic, the boundary integral in the last term of Equation (10) include only the contribution from the contact interface I (rake and flank). The finite-element approximation of the temperature field is DOF t9= 20j(t)hj(x,y), (11) '=1 where DOF is the number of degrees of freedom—which, in the case of the scalar Equation (5), coincides with the number of nodes—and h j- is a bilinear interpolation function associated with the node j. Substituting approximation (11) into the weak form (10) using the Galerkin weighting w(x, y) = h.- (x, y) (12) (i = 1,. . ., DOF), then performing the required integration DOF number of different times, results in the spatially discretized equation 2M1} 37921919; ”2019 +1? +Q.-+ZF ,- 6,, (13) j 1' 1' where the terms of Equation (13) are defined as follows: heat-capacity matrix M ij = Ipé hihjdA , (13a) conduction matrix K”. = Ith" - Vh j(1A , (13b) 18 advection matrix C0. = j pcv - hthjdA, (13c) effective internal forcing Ri = IhiP’work dA , (13d) effective external forcing Q, = Ihiq -nds , (13c) I effective coupling-flux matrix F”. = I 12,.(thj — q)- nds . (13f) I Flux conditions such as convective cooling and flux continuity enter into the matrix Fr , the effect of which is to alter the “stiffness” matrix K ij + Cij. For the case of temperature continuity, the convective-cooling condition is irrelevant, and Fij- vanishes. The effective internal forcing R,- , which accounts for the power generated within the primary shear zone, is evenly distributed to a minimal set of nodes covering the primary shear zone at each stage of the mesh refinement. We will only study the steady state, so that the left-hand sides of Equations (5) and (10) vanish, and the nodal temperatures— 6}- in Equations (11) and (13)—become independent of time t. The (kV h j — q) term in Equation (13f) isolates the contribution of gap heat transfer, as will be shown in a later example (Section 2.4.3). The current algorithm makes use of the background material and source codes of Kwon and Bang [2000, Examples 5.9.1 and 6.6.2]. Alterations to the source codes are made to accommodate Ci]- , Fij , and the source term Ri. To improve accuracy in the present case of advection-dominated flow [Gresho and Lee, 1979] and for the modeling of the primary shear zone, a subroutine is written to incorporate the mesh-creation and mesh-refinement (‘regular’lbisection) algorithms found in Matlab’s “pde toolbox.” 19 2.4. Implementing the interface conditions A matrix equation in the form of Equation (13) for the steady state is written for each of the tool and work regions. Enforcement of interface conditions results in the coupling of the tool equation and the work equation. For a detailed discussion of such coupling, it is felt that the terminology of Equation (13) would unnecessarily obscure the essential ideas. Hence, in the following discussion, we use the notation of a 1D problem, and illustrate how different interfacial behavior can be implemented in the finite-element setting. For this purpose, we refer to Figure 6. The work and tool materials are schematically represented to lie along the x-axis, and they are in contact at the location shared by the work node w2 and the tool node t1: frictional flux Qapplied t1 tool t2 ‘ k A v v v M work w2 Figure 6: Ideal 1-D representation of the work and tool materials A heat flux Q(in W/mz) is applied at the interface. The energy equations governing the work and tool domains are d23 . d6 kwork jfi _ pworkcworkV if“ = 0 9 (14) d29 km] —g’i‘ = o. (15) dx As before, k is the thermal conductivity, p is the density, 6 is the specific heat (per unit mass), v is the cutting speed, Bis the temperature, and x is the distance coordinate. The 20 system is subject to two far-field boundary conditions: aworrlxzxwl = 6m = 0, (16) and 6100],:th = 6a = 0. (17) For simplicity, one element is used to cover each of the work and tool regions. Then, the finite-element discretization of Equations (14) and (15) becomes w w 6 t r 0 [ 11 121 w ]=[Qw1) and [11 121 11)=[Qr1] (18)&(19) W21 W22 9w2 sz t 21 ‘22 9:2 Q12 where, for the work and the tool, respectively, Wij and ti}- (i, j = l, 2) are components of the “stiffness” matrixes, Hwi and 6?t ,- (i = 1, 2) are the temperatures at nodes wi and ti, and Qwi and Q,- (i = 1, 2) are corresponding components of the frictional heat fluxes, being positive for flux inflow into a region. Standard calculations for 1st-order, linear elements show that kwork _ pworkéworkv _ kwork + P workaworkV Lwork 2 Lwork 2 ktool §_ktool tutu)- Ltool Ltool and _ .............. , ............. 2 1 (’21122 _kr001§ ktool ( ) Lroolé Ltool where Lwork = xwz —xw1 and Ltool = xa —xt1 are the element lengths. 21 2.4.1. Uncoupled interface Accounting for the total, effective frictional flux chat is applied simultaneously to nodes w2 and t1, the uncoupled system is modeled by the following combination of Equations (18) and (19): (W11 W12 0 0Y0“) (le\ ( lel W21 W22 0 0 9w2 = Qw2 = flQ (22) 0 0 ’11 ‘12 911 Q11 (1‘13)Q (0 0 ’21 122/0912) (Q12) ( Q12 ) Here, ,6 is the fraction of heat that is partitioned to the work. Being undetermined, ,6 is commonly set to 0.5 to give equal heat partitioning to the work and tool. This is equivalent to applying the two conditions sz + Q11 = constant Q, (23) Qw2 - Q11 = 0- (24) 2.4.2. Temperature continuity The algorithm used here for implementing the temperature-continuity condition follows directly that of Tay et al. [1974]. Under temperature continuity, 9,, =9,, :9, (25) while at the same time, the total amount of applied, frictional flux is sz + Q11 = constant Q. (26) Equations (25) and (26) represent a redundancy in Equation (22) (first equality) that can be collapsed: 22 W11 W12 0 9w1 Qw1 W21 W22 +211 112 9 = Q . (27) 0 t 21 t 22 912 Q12 Hence, the application of temperature continuity to a coincident, contact-node pair reduces the size of the uncoupled system by one [Tay et al., 1974]. In the 2D algorithm used in this paper, the contact nodes on the work and tool boundaries coincide in location to ease the matrix-reduction procedure. The far-field boundary conditions 0m = 6’12 = 0. (28) are applied to Equation (27). Solving for 9 = 0w2 and re-expanding to Equation (22), the heat partitioned to the work element is found to be ,6 = W22 ' 6w2 = [0‘ work [Lwork )+ (pworké5 workV/ 2)] . (29) Q [0‘ work [Lwork )+ (P worke workV/ 2)] + (k tool/Ltool ) 2.4.3. Convective cooling (and flux continuity) Convection-cooling conditions are formulated using the heat transfer coefficient h and the fraction ,6 of heat partitioned to the work as follows: 80 km 3"“ = h(611-6w2)+fl<2, (30) x=xw2 819 ktool[- 7:31] : h(6w2 _6tl) + (1"sz (31) x=xtl Their finite-element weak forms of which are as follows: (M = h (611 —6w2 )+/2’Q, (32) 23 Q11 = h (6.12 - 011) +(1—fl1Q. (33) Substituting Equations (32) and (33) into the first equality of Equation (22) gives: I W11 W12 0 0 Yam) [ Qwr \ W21 (W22 + h) ‘ h 0 6'm _ flQ — . (34) 0 "h (’11 +11) ‘12 911 (l-flh \ 0 0 t21 t22 A912) \ 912 J Referring to Equation (13), the effective coupling-flux matrix in this case is 0 0 0 0 _ _ 0 h - h 0 (Pg) 0 _ 1. h 0 (35) 0 0 0 0 Flux continuity is convective cooling with Q: 0, which leads to a trivial solution if the system boundary is held at ambient, unless a heat source is specified away from the interface. By comparing Equations (22) and (34), it is seen that flux conditions do not reduce the size of the uncoupled system, but rather alters the “stiffness” matrix. The heat partition ,6 becomes a parameter that can be varied. 2.4.4. Mixed interfacial behavior In the 2D implementation of interfacial coupling, the presence of many interface node pairs requires consistent book—keeping. The added complication, however, allows us to consider inhomogeneous interface behavior. 2.5. Simulation results Figure 7 illustrates the refined mesh used for the numerical solution. For imposing coupling conditions, node pairs across the tool-chip interface coincide in 24 location at each step of mesh refinement. A separate refinement near the end of the tool- chip contact helps to resolve the temperature jumps for loose—contact conditions. Table 3 summarizes the parameters used in the simulation. x 10° 2X 10 , _ 14 page 12 chip 1 ,5 E55? E 10 ’E‘ 1. 5:551? v 8 v 3:: .. 8 8 "' en % 4 ‘6 0 VA‘ ‘és'A‘ > V 2 tool > -o 5 0m v 0 A work '1 5‘ 0 5 1 0 1 5 -1 0 1 distance (in) x 10'3 distance (m) x 10“ Figure 7. Mesh used for O-th order problem, and a magnified view 25 Table 3. Parameters used in the 0th—order problem (background temperature) Work piecel: AISI 1045 work piece Tool’: Kennametal TPG322, K-ll Co—cemented Thermal conductivity" kwork : 50.8 N/(s °C) Thermal conductivity‘ ktool : 84 N/(s °C) Density pwork : 7870 kg/m3 Specific heat capacity” Ptool 43V, tool : Specific heat" 511,111,111 = 486 J/(kg °C) 3e6 J/(m3 °C) Uncut-chip thickness’: 173nm Rake angle’: 5 degrees Cut-chip thickness: 519nm Ambient temperature: assumed 25°C Out-of-plane width": 2 mm Cutting speeds“: 2.5 and 4 m/s Calibrated cutting forces,w uncoated carbide tool, 4 m/s: cutting force = 1170N, feed force = 494N. Calibrated cutting forcesf” uncoated carbide tool, 2.5 m/s: cutting force z 1222N, feed force z 516N. (Cutting forces reported' for 4 m/s cutting speed and HfN-coated tool: cutting force = 800N, feed force = 330N) Interface contact lengths . from 0 to 1.2 mm, intimate/temperature—continuity contact; from 1.2 to 1.8 mm, loose/convective-flux contact. 1. i ii ii iii iii 11*. taken from [Subramanian et al., 1993] information on conductivity of carbides is scant; data is taken from CRC Materials Science and Engineering Handbook, CRC Press, Boca Raton, FL, 2001, p. 405 taken from Properties and Selection: Irons, Steels, and High-Perfonnance Alloys, Metals Handbook, Tenth Ed., Vol. 1, ASM International, Materials Park, OH, 1990, pp. 197-198 approximated from Tlusty, Manufacturing Processes and Equipment, Prentice Hall, Upper Saddle River, NJ, 2000, p. 435 forces calibrated from Hfl‘l-coated-tool data to obtain the same average interface temperature reported by Subramanian et al. for uncoated tools (1250°C for 4 m/s; 1100°C for 2.5 m/s) Contact length is reported in Subramanian et al. for the uncoated carbide tool only for the case of cutting at 4 m/s. From Figure 7 of Subramanian et al. [1993, p. 296], the contact length increases approximately linearly from 0.8mm to 1.6mm, corresponding to 3 to 30 seconds of cutting. Since the tool—chip interface temperature for 240 m/min stabilized after approximately 15 seconds, we assume a steady state contact length of 1.2 mm, or 1200 microns. The development of contact length up until the final value of approximately 1800 microns is also taken from Subramanian et al. Because our main interest is in illustrating mass transfer, the same contact length is used for the case of 150 m/min for simplicity. For simplicity, the only coupling between the tool and the chip is through the tool-chip contact (i.e., the flank wear-land is assuming to be zero and conditions are adiabatic there). 2.5.1. Mixed interfacial behavior Two-thirds of the tool-chip contact interface, from 0 to 1.2 mm, is prescribed by temperature-continuity conditions. The remaining, downstream one-third of the interface is convectively cooled with a heat transfer coefficient of h = 1e3 W/(m2 -°C). 26 Figures 8-9 plot comparisons of the heat-partition fields and the corresponding interfacial temperatures. In Figures 8, the heat-partition fields used are compared against what they would have been had temperature continuity been imposed on the entire tool- chip interface: Heat partitioned to work interface, ,6 Heat partitioned to work interface. ,6 for a 2.5«m/s cutting speed for a 4-rn/s cutting speed 2 r . . 2 . - . temperature continuity temperature 00'1”"in fraction of total friction flux fraction of total friction flux 0.51 . I . . 0.51 . / . . 2/3 contact. temperature continuity 2/3 contact. temperature contmurty us contact convection with 0.5 1/3 contact: convection with 0.5 O ‘ heat partition . 0 . heat partition . 0 0.5 1 1.5 0 0.5 1 1.5 Distance from cutting edge x10” Distance from cutting edge x10” (a) along toolvchip interface (m) (b) along tool-chip interface (m) Figures 8a, 0. Comparisons of heat partitions of mixed interfacial conditions to the case of pure temperature continuity Tool-chip-interface temperatures for 66% Tool-chip—interface temperatures for 66% temperature-continuity contact on tool- temperature-continuity contact on tool- 0 C chip interface, with h = 193, [3:05. °C chip interface, with h = 1e3, 13:0.5. Cuttin eed = 2.5 m/s Cutti eed = 4 m/s 1300 9 Sp . - 1600 . "9 Sp . Tool-interfacw Tool-interface/ 1200. temperature 1400' temperature 1100 . 22:22:23? 1 200‘ Chip-interface 1 000 temperature 1000 900 8 00 800- 700 600 1 6000 0:5 1 1:5 40‘)0 0:5 1 1:5 '3 Distance from cutting edge )110'3 along tool-chip interface (m) Distance from cutting edge x10 (a) along tool-chip interface (m) (b) Figures 9a, b. Interface temperatures corresponding to Figures 8 27 Figures 9 illustrate how the chip—interface temperatures overtake the tool interface temperatures near the end of the tool-chip interface, where the heat partition for mixed conditions overtakes the heat partition for temperature continuity.5 The size of the temperature jumps also correlates with the differences in the heat partition fields. The location of the crossover point in each set of data, however, is not easy to predict. 2.5.2. Temperature fields Finally, Figures 10a, b plot the full solution to the O-th order problem corresponding to the two cutting speeds. Temperature field for 2.5-m/s cutting speed Temperature field for 4-m/s cutting speed r “. a Wé‘\’\§ %&)¢§V§ "v“ " i I‘lllll‘n. . l r 'lgl'w":‘i ‘ I, @333 'gz‘VAVAVAVAv -1 -0.5 0 0.5 1 1.5 x axis (m) x 104 x axis (m) Figures 10a, b. Temperature fields (see also Figures 9a, b) The extremely rapid rise of the temperature field across the primary shear zone, and a less rapid but still substantial rise adjacent to the tool-chip interface, both contribute to the need for mesh refinement. 5 The crossover near 1 mm in Figure 8b is considered an imperfect attempt of the numerical scheme to satisfy temperature continuity from 0-1.2 mm and to shift to convective conditions. The main crossover points of interest are associated with the constant heat partition of 0.5, as this is the cause of the difference between Figures 8a and 8b. Considering that in the case of mixed boundary conditions, the numerical scheme receives no error feedback in its attempt to satisfy temperature continuity, the match in Figures 8 over portions of temperature continuity is very good. 28 2.6. Summary of the temperature problem The nature of heat transfer across the tool-chip interface has been studied for the uncoupled interface, the temperature-continuity interface, and the convectively cooled interface. It has been found that the nature of interfacial temperature jumps in all cases except temperature continuity is directly related to the interfacial heat partition. Further, if the work’s heat partition is higher (lower) than the value that it assumes in the case of temperature continuity, then the work’s interfacial temperature is also higher (lower) than the tool’s interfacial temperature, although the crossover points in the two cases are different in general. Namely, the presence of crossovers on the ,6 plots (e.g., Figures 8) correlates with the presence of crossovers in the interface-temperature fields (e.g., Figures 9). Hence, interfacial temperature jumps are directly related to the heat-partition function in the 1D case (see Appendix), and can be numerically correlated to the behavior of the heat-partition function in the 2D case. The missing link to the actual nature of contact would be a relationship between the heat-partition function and the mechanical nature of frictional contact, but this is beyond the reach of the present work. In the next chapter, the interfacial temperature jumps found in this chapter are assumed, and their effects on dissolution wear will be studied. 29 Chapter 3 THE FIRST-ORDER PROBLEM OF SPECIES TRANSFER To completely define the first-order problem, it is necessary to specify the species number N and how mass is exchanged. It is proposed here that the equations of the Frank-Tumbull mechanism be used. The Frank-Tumbull mechanism describes the combination of an interstitial-A atom and a B-lattice vacancy to form a substitutional-A impurity, or A, + BV —> AS. (36) This reaction is illustrated in Figure l 1: interstitial substitutional W0 0 o o 0/o\o/ o o 0 Ci) 0 e o o codex oooo vacancy Figure 11: Schematic of the Frank-Tumbull reaction 3.1. The time-dependent equations of the Frank-Tumbull reaction and mass transfer The governing equations are written using Equations (0) and (p) of Table 2 (with k = ‘i’, ‘V’, and ‘s’) for the Reaction (36), where references to A impurity atoms and the B lattice are understood, and all terms are defined in the “Key to Variables, Parameters, and Operators” on page ix: 3O -a—-X—i-=V(Di VXi)-VXiV*—£ (37) a: c BXV F FV psz -—=V'(Dv VXV)—VXV -v*-—i+—’—— (38) at c c 8X. =_VXS -v*+£ (39) at c F X', XV, ‘—s=kfwd[XiXV_ ”‘1 6“ X8] (40) C Xs,cq where Fs lc embodies the rate of production of substitutional impurities via the Frank- Tumbull reaction [Sturge, 1959; Giese etal., 2000]. The term dem / c is defined by F V, psz c =Ic\,(XVdaq -XV) (41) and represents the generation of vacancies within the primary shear zone only (see Figure 3). The impurity subscripts ‘i’ and ‘s’ can refer to either tungsten or cobalt. In effect, each diffusing tool element is modeled separately as a dilute tracer component. Additional assumptions are the constancy of concentration c, the diluteness X k 2m Fm I = I X k (— Fs + FV, psz)| ), and the neglect of the approximation (lel » diffusive mobility of substitutionals as compared to the mobility of interstitials and vacancies (DS « Di,DV ). Note that no equation is written for the chip species (iron) in its role as the substrate lattice; however, its presence implies the approximate constancy of the total concentration c, which is an important simplification. Solubility data from binary phase 31 diagrams are also reported relative to the chip species. Lastly, the concentration- averaged velocity v* is approximated to be that of the chip advection velocity. 3.1.1. Boundary conditions for mass transfer The decomposed tool components are postulated to enter the chip lattice as interstitial impurities. Then, dissolution is controlled by the flux of interstitials entering the chip boundary (viz., the crossing of the bold line from tool to chip in Figure 3): 3X- Di 3;? = [disskdiss (Xi, eq — Xi ). (42) where n is the outward normal distance from the tool interface into the depth of the chip, ldiss is a length that is used to render a volumetric rate into an area flux, and kdiss is the rate constant for the dissolution of tool material into chip, defined by AEdecomp ( A? ' j k ' = ex - —_ ex "' __a_Cth_n_ . 43 diss AOVFe P[ R 0 P R 0 ( ) Here, A0 is a calibration constant. According to Shewmon [1989, p. 70, pp. 74-75], given the diluteness assumption, exp(— Agdccomp / (R0 » can be interpreted as the mole fraction of decomposed tool components under chemical equilibrium (i.e., the equality of Condition (v) in Table 2), and Vpe exp(— Agacuvn / (R 0)) can be interpreted as the jump frequency of tool constituents moving into the work lattice under chemical equilibrium.6 6 The interpretation makes use of the idea of activated complexes (i.e., vibration-activated openings on the chip surface through which tool constituents pass). In an earlier edition of his book, Shewmon [1963, p. 60] notes that the “precise definition of v is one of the more difficult aspects of a rigorous theory. However, it is usually taken equal to the Debye frequency.” Shewmon [ibid., p. 60] continues to say that the entire derivation hinges on the assumption that Agacm “is a state function, that is, that the free energy of the activated complex attained by the reversible process is the same as the free energy of the 32 In this interpretation, we can also view the interstitial boundary flux Di 3X i / an as the product of the concentration of open sites for interstitials, times the product of the probabilities (i.e., mole fractions) of decomposition and activation, times the frequency that the tool constituents are allowed to jump into a lattice site within the chip. Hence, we define dissolution wear as the combined sequence of events of decomposition at the interface and the subsequent advection-diffusion-reaction within the chip region. X V and X s are simply given zero mass flux boundary conditions; the validity of this choice for the advection boundaries is discussed in [Zienkiewicz and Taylor, 2000b, pp. 44-46]. The same condition applies for X i along all parts of the boundary except the tool-chip interface. 3.2. Connection to previous work by Kramer The role of the rate constant is to determine the rate of rise of the interstitial distributions but not its saturation value X i, eq . In fact, simulation results using Equation (42) shows that the solubility X i, eq is the maximum value of X i that is achievable along the contact interface (Section 3.5.3). This is consistent with the earlier works of Kramer. In particular, mass transfer at the contact interface is rate-controlled by the supply of interstitials, and the solubility X i, eq governs the maximum amount of activated complex that really occurs in nature. This would be true if the complex, like a vacancy, existed for a period long enough to let the surrounding lattice completely adjust to the presence of the complex. . .. but if the system is not in thermal equilibrium (present author’s note: i.e., chemical equilibrium), the entire procedure may become only an approximation of doubtful value... One of the main justifications of this analysis is that it works. It is not clear to many authors just how the assumption discussed above can be fulfilled, but whenever the predictions of this theory have been experimentally checked, the theory has been vindicated.” 33 dissolution wear. At high temperatures, the value of the impurity distribution at the interface approaches its solubility value, and the dissolution effect approaches its peak. 3.3. The time-independent equations To model time-independent behavior, we set both B(')/at = 0 and Fs = 0 in Equations (37)-(40). It is implicitly assumed that the chronological order of occurrence of (i) the achievement of a macroscopic steady state, and (ii) the local saturation of the Frank-Tumbull reaction, is immaterial. Under these conditions, X s can be eliminated by recasting the equations as follows: 0=§¥i+§§=v.(oivxi)—V(xi+Xs)-v* (44) a: a: F 0=§£l+3X—S=V-(DVVXV)-V(XV+XS)-v*+-l’-P-s—z- (45) a: a: c XS,” WhCI‘C XS = KequXV = XiXV (46) Xi eqXV,eq and K eq = X 5,6,] / (X i,eqX V,eq) is the equilibrium constant. Equations of the type (44) and (45) have been referred to by Sturge [1959, p. 298] as the equations of atom conservation and lattice-site conservation. 3.4. The difference between W and Co profiles; non-dissolution mechanisms The difference in W and Co concentrations in Figure l is not likely due to electrical mechanisms. WC-Co is metallically bonded, and when it is dissolved into Fe, the components W and Co are most likely charge-neutral interstitials. 34 Comparisons of atomic radii and decomposition energies do not account for the difference in the W and Co profiles, either. Since Co and W have similar atomic radii as iron, the activation energy required for Co and W to enter the chip lattice (Agacfivn ) should be similar. In addition, the free energies of decomposition entering the calculations for W and Co are one and the same (viz., Agdecomp of WC —) W + C). Referring to the impurity-flux condition, Equation (42), the only remaining factor for explaining the difference between W and Co profiles is the solubility X i, eq . While the solubility of Coin Fe is much higher than the solubility of W in Fe [Hansen et al., 1958, p. 472], the actual amount of Co dissolved in the chip is capped by the supply of decomposed tool material, and hence should be in the same molar proportion to W as found in the tool. To check the W-Co ratios before and after cutting, the data of Subramanian et a] [1993] is converted into molar bases in Figures 12a, b: Original data: mole fraction of w in chip Original data: mole fraction of Co in chip _‘ J; 0-8 240 rn/min 240 m/min 0.6 mole fraction ( mol/(100 mol of chip) ) mole fraction ( mol/(100 mol of chip) ) 0.4 150 m/min 0.2 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4 depth into chip (pm) depth into chip (pm) (a) (b) Figures 12a, b. Experimental concentration profiles, deduced from Subramanian et al. [1993]. Empty and filled circles denote data points. in (b), the 240-m/min curve asymptotes to the same non-zero value as the 150-m/min curve. 35 Incidentally, the maxima and minima of the concentration profiles for the W and Co for a given speed occur at the same depths, which shows that Co has similar mobility as W as both constituents move through the chip lattice. We now check the molar ratio of W and Co within the chip and compare it to the corresponding molar ratio within the fresh tool material. The areas under both the W and the Co profiles are assumed to be representative of the total moles per unit length in the chip flow direction. For the case of 240-m/min cutting speed, we find that the W-to—Co ratio of absorbed moles per unit length (i.e., the ratio of the W and Co areas for the case of 240-m/min cutting speed) is 5.34-to-l. On the other hand, it turns out that the molar/atomic composition of the tool is 45.6% W, 45.6% C, and 8.1% Co (with a trace of TaC-NbC), or a W-to-Co ratio of 5.63- to-l. Considering that the 5.34-to-l ratio is calculated from a single sectioned profile, it is considered a good match to the 5.63-to-1 ratio. Hence, we find consistency with the hypothesis that Co maintains its molar ratio with W before and after cutting. For the case of lSO-m/min cutting speed, the same process of calculating the W-to-Co molar ratio per unit length from Figures 12a, b returns a low value of 2. The cause of this low value is the non-vanishing tail in the Co profile after a depth of 0.1um in Figure 12b. We choose to ignore this tail on the hypothesis that the detected Co traces are related to the W traces in roughly the same molar ratio inherited from the fresh tool, and the fact that the W profiles decay to zero.7 Using maximum values from the profiles instead, we obtain a W- to-Co molar ratio per unit length of 6.25. This ratio is consistent with an increased amount of interface wear (e. g., abrasion) not accountable by dissolution, and it is 7 The other interpretation would be that the W profile decays asymptotically to a small, but nonzero value, but this is judged less likely because such a non-vanishing tail with zero flux would require us to introduce a condition of mass buildup at some pre-determined depth. This issue will be revisited later. 36 explained as follows. Suppose that, for every 100 moles of tool constituents decomposed (i.e., prior to diffusion and advection), 0.96 moles each of W and Co are removed due to various interactions at the interface, and are made inaccessible to diffusion into the chip. Then the W-to-Co molar ratio would change from 45.6-to-8.1 prior to cutting, to (45.6 — 0.96)-to-(8.1 — 0.96) as detected within the chip afterwards, or equivalently, 6.25-to—1. In Section 3.5.4, this calculation is revised to make use of the concentration fields that have been extrapolated from experimental data using the Frank-Tumbull interpretation. To summarize, our interpretation of the profile data of Subramanian et al. indicates that the W-Co molar ratio found in the chip reflects, in the main, the W-Co molar ratio found the tool. Differences in the W-Co molar ratios before and after cutting can be used to gauge the amount of worn cobalt that is not attributed to dissolution wear. Among non-dissolution-based wear mechanisms are the deposition of tool constituents onto and subsequent advection by the chip (e. g., asperity exchange favored by surface energy changes [Rabinowicz, 1965], 2-body abrasion and deposition through applied pressure, plowing [Suh, 1986], etc), the sweeping action by loose debris in rolling friction (i.e., 3-body abrasion [Rabinowicz et al., 1961; Rabinowicz, 1977]), the deposition of tool constituents onto rolling debris at high pressure—with the number of mechanisms likely limited only by the imagination. 3.5. Results of the species-transfer problem The interfacial temperature field is gotten from the analysis of Chapter 2, and the result for the 4-m/s cutting speed is reproduced in Figure 13. It is seen that the interfacial 37 temperatures can vary significantly from its average value. The temperature jump in the region near the end of contact is due to a hypothesized region of loose contact. 'C Extension of Average Interfaclal Temperature Data 1600 ' ' tool average: 1265’C \_ 1 400 chip average: 1235‘C \ Subramanian et al.: 1250'C 12007' 1 000 > 800* 600 0 as 1 15 Distance along interface from cutting edge (m) x 10{3 Figure 13. Variation of temperatures away from the average Figure 14 shows the region of interest and the numerical mesh used for the species-transfer problem: depth into chip (m) 0 1 2 3 distance along flow direction (m) X 106 Figure 14. Mesh used in the 1st-order, species-transfer probelm The x axis from 0 to 1.8 mm corresponds to the tool-chip interface parallel to the flow direction where interstitial impurities are introduced. The y axis corresponds to a very small portion of the primary shear zone where vacancies are generated. The y direction is 38 the chip-depth direction along which concentration profiles vary, and is plotted on a highly magnified scale. The depth distances are so small that the temperature, and hence the material properties, only vary in the x direction. A product of interstitials and vacancies leads to the humped concentration profile that is the basis of the present application of the Frank-Tumbull mechanism; for this purpose, a special vacancy distribution is chosen. An example of one such distribution, plotted topographically on the mesh of Figure 14, is shown in Figure 15a: Steady-state X i (W-lnterstitial) concentration in iron, 4-m/s cutting speed Steady-state Xv concentration in iron. 4-m/s cutting speed .0 0 00 l I I I 2 E ‘5 . 2 TE .- .‘\\\“ :‘ E a :0-02- .101 ~ ‘ *1 C ,s‘.\\l\11\\\“mg. 1 V _ haw , ‘1 v .5 0.01 1 1 “ : L3 .2 0 . ‘ - g 4 3 g E 7 "’" 1 E -7 4 x 10 depth 0 0 x 10 depth 0 flow direction (m) x 10 . ' flow direction (m) . . dlrectron (m) direction (rn) (a) 02) Figures 15a, b. Sample XV and X; distributions The rise of X V in the depth-direction (y—direction) is obtained empirically by targeting the iterated solutions of the Frank—Tumbull nonlinear equations (Equations (44)-(46)) towards a match with the rise of the profiles found by Subramanian et al., and by retroactively adjusting the input supply of vacancies within the primary shear zone (Equation 7) until the match is within tolerance. The iterated steady-state solution, based on the method of quasi-linearization [Kwon and Bang, 2000, p. 539], is set up so that the 39 value of X S after the first iteration is a simple product of X i and X V ; subsequent iterations account for species transfer. In Equation (41), kv is set to be a hyperbolic tangent function of y that is used to introduce a quick rise of vacancies within the primary shear zone, and X V, eq controls the height of the vacancy distribution. The form of vacancy generation in Equation (41) is numerically less demanding than a Dirichlet condition, and it also has the property that as the advection speed increases, the height of the vacancy distribution decreases. The underlying assumption is that the amount of vacancies generated is bounded, and the height of the plateau generated by advection is determined by a competition between the vacancy-generation rate and the rate at which vacancies are carried away. A smooth rise in the x direction is also introduced to avoid the need for a highly dense mesh. Figure 15b shows the interstitial distribution (solution to Equation (44)) with which the vacancy distribution interacts. For convenience, the parameters governing the boundary condition of the interstitial distribution are collected in Table 4. Before finishing the discussion on the interstitial distribution which leads to the main results, a digression is taken to explain the two most important parameters, namely, diffusivity and solubility. 40 Table 4. Parameters used in the lst-order problem (mass transfer) 3X - Agdecom A- ' M355 flux Dr a_nl= Idiss [ AOVFe CXP['—'7z'E—E‘ exp Tiféfl '(Xi.cq — Xi) A0: I (assumed; as long as this number is of order 1, the sensitivity of the steady-state results to its variations is low) ldiss : 2.5 X 10—8 m (2: lattice distance; D. Gaskell, Introduction to Metallurgical Thermodynamics, 2nd ed., Hemisphere Publishing Corporation, 1981, pp. 117-118) VFe : 1X 10+13s-l (z Debye frequency; D. Gaskell, Introduction to Metallurgical Thermodynamics, 2nd ed., Hemisphere Publishing Corporation, 1981, p. 118) Agdccomp : 9000 cal/mol (from the reaction WC—)W+C; O. Kubaschewski, E. L. Evans, and C. B. Alcock, Metallurgical thermochemistry, 4th edition, Pergamon Press, Oxford, 1967, p. 429. Actually, it is 9000—0.4t9cal/mol, but exp(— (9000 - 0.46)/(Rt9)) becomes exp(+0.4)- exp(- 9000 / (R 6)) , and the exp(+0.4) term is grouped into the constant A0 .) Macaw, : 18000 cal/mol (4-m/s cutting speed); 21000 cal/mol (2.5-m/s cutting speed) (curve-fit value that controls the maximum of X s (adjusted after abelow is determined» X i, gq = aX W, eq and X 5, eq = (1— a)X W, eq , where a = 0.0375 (4-rn/s cutting speed), and a = 0.0150 (2.5-m/s cutting speed) (ais the first curve-fit value determined) Xw’ eq = 9x10-11TC3-1.23x10_7T62. + 7.6X10—5 Tc -9.8)<10”3 , where TC is temperature in Celsius. (Value of X W, eq taken from solubility curve of W-Fe binary phase diagram, p. 734 of Constitution of Binary Alloys, 2nd edition. Metallurgy and Metallurgical Engineering Series, McGraw-Hill, New York, 1958. Cubic best fit from Microsoft Excel) X Co, eq = a} X W, eq : at: 0.125 (4-m/s cutting speed); at: 0.185 (2.5-m/s cutting speed) (§3.5.4) Remaining parameters: X V, cq = 0.035 (curve-fit value that affects the rise of X 5 through its influence on X V ’5 rise) Di : 1x10"10'925 m2/s (4 m/s-cutting speed); 1x 10’11'725 m2/s (2.5-mls cutting speed) (curve-fit value that controls the spread of X s (adjusted after ahas been determined» Dv : 1x10"14 m2/s (curve-fit value that prevents X V from losing its shape due to advection) Ds : neglected in this analysis kv : quick rise to a value of 0.25 (curve-fit function that controls the rise of X V , and hence X s , within the primary shear zone) 0+5 s' kfwd : 1x1 1 (curve-fit value that controls the Frank-Turnbull reaction rate during relaxation) p0 : 0.3 x 10+9 N/m2 (force/contact area; needed for studying effect on diffusivities, both in decreases in magnitude and variations across the ferrite-austenite transition temperature) 3.5.1. Unusually low values of mass diffusivities 41 Figures 16a, b show a comparison of the tungsten-interstitial diffusivity DW-in-Fe and iron-vacancy diffusivity DV used in the simulations versus those found from tablessz Comparison of diffusivities in Comparison of diffusivities in iron, 2.5-m/s cutting speed iron, 4-m/s cutting speed c -u . . . .4 - . . A DW-in—F93 table-lockup : own“: table-lockup A $ ‘6 > \ & -8> . NE W tablelookup "E by: table-lockup .. s -8» ‘< -10 2: .E‘ 22 j .2 g -10* s .12 -—- — -*——- a ~ —-—-—— -~—- % Dw-in-;:e: used in simulation 3 -12 DW-in-Fe3 used in simulation . V 9 0244* " —\ 3 -14 —°— DVI used in SimUIBIion \ 0V: used in simulation '1"0 075 1 115 4‘0 015 1 115 distance from cutting edge x 10'3 distance from cutting edge x 10'3 along tool-chip interface (m) along tool-chip interface (m) (a) (bl Figures 16a, b. A comparison of available values of diffusivities versus actual values used in simlations for (a) 2.5-m/s cutting speed, (b) 4-m/s cutting speed A small DV prevents the vacancy distribution from losing its shape downstream of the primary shear zone, which is needed to model the quick rise (within a 0.1-um distance) of the concentration profile of Subramanian et al. [1993] (Figures 1 and 12 in this work). Also, Di must be small enough to force a quick decay (also within a distance of 0.1 pm). Values of diffusivities available in the literature are simply too large to achieve these 8 Spatial dependence of mass diffusivity is due to the background temperature and pressure fields from the 0th-order problem. The sharp drops in mass diffusivities near the origin correspond to the ferrite-to- austenite transition point, which, at a pressure of 0.33e9 Pa, is approximately 718°C. Table-lookup values are taken from Elliot et al. [1963] as Arrhenius functions of the absolute temperature. Constant values of mass diffusivities are adopted in the simulation for a practical reason: a sharp drop in diffusivity requires mesh refinement for resolution and numerical stability, especially in the presence of strong advection. Physically, however, most tool wear occurs downstream of the cutting edge (e.g., the boundary values of interstitials in Figure 15b) where the mass diffusivity is almost constant, hence the use of constants. As mentioned in Section 3.4 (discussion of Figures 12), the diffusivity of Co is rate-limited to that of W. 42 effects. In fact, the use of the table-lockup diffusivities leads to vacancy and interstitial distributions that are shown Figures 17a, b: Xv concentration In iron using table- Tungsten-Xi concentration in iron using IOOkUp diffusivity, 2.5-m/s cutting speed table-lockup drffusivrty, 2.5-W8 cutting mole fraction ( moV(1 mol of chip) ) mole fraction ( moV(1 mol of chip) ) Figures 17a, b. X V and Xi distributions using table-lockup values of Figures 16 In contrast to the distributions of Figures 15, neither a rapid—rise function nor a decaying function is present, and the product of the distributions in Figures 17a, b will not result in a humped distribution in the y-direction as required. Pressure and damage are two possible reasons for the seeming prevalence of low values of diffusivities. Although pressure does not enter explicitly in the equilibrium condition (equality of Condition (v) in Table 2), it tends to decrease the values of diffusivities. An expression by Shewmon [1989, pp. 84-85] illustrates this relationship: 36 1n 2 D ____‘7actiRy;tion , (47) p0 ldiss VFe 6fixed where activation is the activation volume (in m3/mol), and it is known to be smaller than the molar volume of the lattice metal (viz., iron). Assuming constancy of all terms except D for simplicity, we have that 43 D (P0) = exp[— Foil—activation J (48) D(1 atmosphere) R 6 For a pressure of p0 of 0.33 X109 N / m2 at 1000K and a maximum fir-activation value of M Fe + PFe z 7 x1076m3 / mol , we arrive at a maximum drop in diffusivity by about a factor of 3. Damage mechanisms may be needed to achieve the levels of reduction in diffusivities as suggested by Figure 16. That is, the concentration profiles from experiments may reflect shear damage so severe that the mobility of impurities is significantly impaired. However, no simple, independent test of this hypothesis is available at this time, and the mass diffusivities are considered curve-fit parameters. 3.5.2. The determination of interstitial solubility; post-machining relaxation In an early development of the model for tungsten impurities, X i, eq and X 5, eq are both set equal to X W, eq because the fractions of interstitials and substitutionals among a given amount of impurities are unknown. Figure 18 shows the result of a simulation for the steady-state, substitutional distribution under the assumption X i, cq = X W, eq = X S, eq for the 2.5-m/s cutting speed. The section taken—where comparison is made with the profile from the data of Subramanian et al.—corresponds to the chosen distance of x = 1.8 mm; this is the end of the tool-chip contact where impurity concentrations are near or at their maximum for a variety of loading conditions. Comparison of w profiles in x 104 iron, 2.5-m/s cutting speed solid: experiment dash: computed )6», mole fraction ( moV(1 mol of chip) ) 1 2 a 4 depth into chip (m) x 10'7 Figure 18. Good match in an early run that turns out wrong Despite a good match, it is remembered that the measured concentration also includes the residual interstitial distribution at the end of cutting (i.e. steady state). The residual X i field is given as a topographical plot in Figure 19a, and its sectioned profile is superposed onto Figure 18 and presented in Figure 19b: Steady-state X i (W-lnterstitial) Comparison of W profiles in iron at concentration in iron, 2.5-m/s 10‘3 x = 1.8 mm, 2.5-m/s cutting speed x a U steady-state X i (or residual interstitials after machining) solid: experimental profile of W in Fe dash: steady-state X5 mole fraction ( moV(1 mol of chip) ) ‘ \\\\\\\\\\\\\\\\ .s, it“ I \\\\ \\\\\\\\\\\\\\\ 1 \ \\\\‘& x u. R \\ \\\\\\ \\\\\\\\\ \\\‘x\‘ _ \ \\\\\\\\\ “\- \\:§:§§‘\‘\e¢s\e it‘sxxsx‘ssxsgt \\\\ _ \‘ ., \\‘\:\\\\\‘\\§‘C\ .. w, a / Ill/l, mole fraction ( moV(1 mol of chip) ) -1 1 . L 2 3 4 depth into chip (m) x 10'7 Figures 19a, b. Why the good match of Figure 18 does not work The sum of the steady—state X s and X i distributions no longer agrees with data. Evidently, X i, eq , which controls the maximum of X i along the tool-chip interface, must be decreased, as reflected by the multiplicative factor ain Table 4. In 45 fact, the value of afrom the curve fit represents the fraction of impurities that can be identified as interstitial impurities, as seen from the following equations: qu = axwflq and ngq =(1—a)Xw,eq (49) Figure 20a show that decreasing X i, eq forces the steady-state X 8 distribution to decay too rapidly. A completely analogous situation occurs for the case of 4 rnls cutting speed: Comparison of W profiles in iron at Comparison of W profiles in iron at x 104 x = 1.8 mm, 2.5-m/s cutting speed x 106 x = 1.8 mm, 4-m/s cutting speed N O steady-state Xi steady-state x; d 9' .5 9 experimental profile of i /WinFe steady-state X3 experimental profile of / W in Fe . steady-state X3 mole fraction ( moV(1 mol of chip) ) mole fraction ( mol/(1 mol of chip) ) A db I (.11 1 a {3 4 1 a a 4 depth into chip (m) x 10‘7 depth into chip (m) x 10'7 (a) (b) Figures 20a, b. Attempts at matching experimental W-concentration profiles using only steady-state W distributions from the model Increasing Di delays the decay of X s in Figures 20, but the spread of the distribution is unacceptably increased (not shown in Figures 20) beyond the experimental-data curve. At this point it is thought that a relaxation process may take place after cutting— but before quenching—such that the tail of the steady profile would ‘fill up’ to match the experimental curve. To test this hypothesis, we model Equations (37)-(39) with time dependence, zero advection speed, homogeneous flux boundary conditions, and with the X i , X V , X S distributions inherited from the steady-state calculation. Also, the temperature field is preserved as quenching has not yet occurred. The result of such a 46 relaxation run for the case of 4—m/s cutting speed is shown in Figure 21a. The solid X 5 curves correspond to the relaxation times of 0, 0.9, and 3.6 milliseconds; it is found that the peak of X s rises to a maximum value at approximately 0.6 milliseconds before it decreases thereafter. For a choice of a uniform and steady reaction-rate constant kfwd , the match to experimental data is considered reasonably good.9 Figure 21b shows the relaxed X 5 field at the final time; a practical choice of the coarsest mesh possible leads to Wiggles on order of 3% of the maximum value of X s : Xs-relaxation profiles of tungsten in Relaxed Xs concentration in iron A 4 AX 10-3 iron after cutting at 4~m/s cutting after cutting at 4-m/s cutting speed A Iu r » — _ _ \ . . o. , . , E 2 2 [I 4I_ _ _ ‘/ O 8 1 . . Q I / 3 dash: experimental data E 3 k E: 6 1: no relaxation is X 10' I ,_ 2: 0.9 millisec. relaxation —° 8 I L“ Ar ‘5 . . - - E , ,. , g 4 3.3.6 millisec. relaxation . r, 6 I, :1 / v g 4 , .,;/,/,/// ‘9 s 2 ’ 7: o 1 "5 o _ r0 2 4 . . . ‘5 ‘0 1 2 3 a .7 E depth into chip (m) x 10 (a) (b) Figures 21. (a) Relaxation of X5 after cutting; (b) Xs after 3.6 millisecs. of relaxation Figure 22a shows the final outcome of the relaxation simulation (starting from the state of Figure 20b) at 3.6 milliseconds. Figure 22b plots the volume under the X s plot as a 9 The use of a non-uniform kfwd can improve match but extra physical mechanisms have to be introduced. The size of kfwd determines the speed of relaxation; for example, the elapse time for the relaxation process to nearly exhaust the tungsten X i in Figures 19b is roughly 3.6 milliseconds, corresponding to kfwd = 1e5 s". The uniformity of kfwd implies uniformity in reaction rates given a uniform driving force; however, because a low value of DV maintains a vacancy—depletion zone near the interface, there are no renewed, Frank-Turnbull reactions there. Consequently, the rise of the X s hump is not significantly altered during post-machining relaxation. The elapsed time after machining but before quenching of the chips is not given in the original data. 47 function of relaxation time after cutting. It is seen that the residual X i field is essentially exhausted through its interaction with X V . Comparison 0‘ W profiles in iron at Evolution of volume under tungsten x = 1.8 mm, for a relaxation time of x8 plot during relaxation and after A x 1 04; 3.6 millisem‘atter 4—mlsvcuttlng 1 4x 10.12 imam at a speed of 4 ntls E o ‘5 8- experimental profile :5 of W in Fe "E 1.3. E e» .5 T: relaxed X3 ‘8: 1.2. E t v 4» 2 5.5 , E o relaxed X . 1.1 s 2' E (E) G A 1 . . . 0 1 2 3 4 0 1 2 3 depth into chip (m) x 10'7 relaxation time (S) x 10'3 (a) (b) Figures 22a, b. A check on the exhaustion of tungsten X i during relaxation 3.5.3. Solubility as the maximum amount of dissolution wear achievable Figures 23a, b show how the equilibrium solubility X i, eq (as defined by Table 4) places bounds on the amount of a tool constituent that is made available for dissolution in the chip.10 Plots are shown for tungsten; since X i, eq for cobalt is some fraction of X i, eq for tungsten (viz., the product of aand (p as discussed in Sections 3.5.2 and 3.5.4), the plots for cobalt are exactly analogous. If Agactivn is very large within the expression for the rate constant of dissolution kdiss (the second Arrhenius exponential in Equation to The wiggles in X i near the origin reflects the difficulty with which the coarse mesh tries to resolve a sudden increase. Fortunately, these wiggles do not pose a significant problem in the convergence of numerical solutions, but they do affect accuracy. The tradeoff in mesh coarseness is necessary, as the relaxation runs cost well over 8 hours to run on a computer with a 1.2 GHz processor. 48 (43)), then the spatial increase of X i towards its maximum amount is delayed; as agacfivn increases indefinitely, X i tends uniformly to zero because the boundary flux does. Comparisonotboundarngand Comparisonoiboundarngand X i. sq (tungsten) at a cutting X i. .q (tungsten) at a cutting 12x104r speedot2.5 ml: T 5x10: speed ot4mls .5 9 .3 .3 '5 ‘5 ”5 "6 3 > 4 _ E 8 3 3» l Q 6~ , E 8 E 2 V 4 . V C C 1 _ 3g 2. steady-state X; at % steady-state X; at g interface boundary g 0 interface boundary 2 O 2 '2 a i 23 e». '1 a a é 5 distance in the flow direction (m) x 10° distance in the flow direction (m) x 10° (8) (b) Figures 23a, b. A check of Kramer's hypothesis that dissolution wear is bounded by solubility It is seen that in going from 2.5 m/s to 4 m/s, the maximum of the boundary supply of interstitials increases by 40 times. While the Arrhenius dependence on temperature of the boundary interstitial flux is influential (Equations (42) and (43)), Figures 23a, b show that solubility is the controlling factor. Further, the sudden change in slope at 1.2 mm reflects the drop in the interfacial temperature of the chip in relation to the tool in the last thirds of contact (see Figure 13). 3.5.4. Amount of worn cobalt not due to dissolution wear By comparing the molar ratios of W to Co within the chip and within the tool, the amount of worn cobalt not due to dissolution wear can be estimated. Relaxation of the 49 tungsten profile after cutting at 4 m/s has already been shown in Figure 22b. Figure 24a shows snapshots of the relaxation process for cobalt impurity atoms after cutting at 4 m/s, and Figure 24b is a check on the exhaustion of X i -cobalt during relaxation: Evolution of volume under cobalt X3 - i m ' Xs re axation pro es of cobalt in plot during relaxation and after 4 iron after cutting at 4-rn/s cutting A x10 1M1043 . Cutting‘ataspeedot-tm/s' .9 15 ' 6 i — 2 ‘5 1 dash: experimental data 1,3- TE) 10» 1: no relaxation ‘ NE 5 2: 0.7 millisec. relaxation .5 1.7. E 3: 4 millisec. relaxation :3 (U E 5» E 1.6» E \- 22 l 15 2 O» 1 O 0 1 2 3 4 1 40 1 2 3 4 distance into chip (m) x 10" relamtion time (s) x 10‘3 (a) (b) Figures 243, b. The relaxation process. A check on the exhaustion of cobalt Xi during relaxation Concentration data suggests that the mobility of Co is similar to the mobility of W as tool constituents diffuse into the chip lattice (Section 3.4). For simplicity, then, the mass diffusivities of Co and W are set equal. The solubility of Co is source-limited to the amount that is originally present in the tool; in particular, the original W-Co molar ratio within the tool is approximately preserved. In the simulations, X Co, cq is set equal to ¢X W, eq , with (oas given in Table 4.1' Hence, with the exception of w , all parameters in the W and Co simulations are the same. Given these assumptions, the agreement is deemed reasonable. The thick tail of the experimental curve has been discussed in H That 47 has to be empirically determined may be a reflection that not all Co particles enter the chip. 50 Section 3.4 (Figure 12b and discussion).‘2 The initial dip in the experimental data of Figure 24a (dashed curve) may represent residual X i -cobalt due to quick quenching; however, no independent confirmation is available at this point. The quotient of the integrated X s ~tungsten evolution in Figure 22b and the integrated X s ~cobalt evolution in Figure 24b is taken to be the evolution of the total W- Co mole fraction within the chip. This W-Co mole-fraction evolution, corresponding to the case of 4-m/s cutting speed, is shown in Figure 25a: Evolution of W-Co mole fraction Evolution of W-Co mole fraction during relaxation (after cutting at during relaxation (after cutting at a speed of 4 m/s) a speed of 2.5 m/s) 2 7.35 . . a 2 5.2 - fl 8 8 .5 1 .2. E . z 725* J z 5'15 E 3 3 mi - g E, 5.1 . l v 7.151 v s a .. i 'o-a ’3 7" g 5.05- 4 1.3 7.05i 3 o o E 7 . . . E 5 - . . . 0 1 2 3 O 1 2 3 4 5 relaxation time (s) x 10“ relaxation time (s) x10"i (a) 0)) Figures 25a, b. The W-Co ratio within the chip as residual X i ’s are exhausted during relaxation tor (a) 4-m/s cutting speed, (b) 2.5-m/s cutting speed. The X i of both species approaches exhaustion when the amount of X 8 stops increasing; hence, we take the final W-Co mole fraction within the chip to be approximately 7.35-to- 1 as indicated by Figure 25a. We note that the molar composition of the tool is 45.6% W, 12 That is, we disregard the thick tail as an artifact because non-zero asymptotes at large depths do not reconcile with the usual picture of exponential decay. If need be, the thick tail can be modeled either by a mass buildup at some depth y, or by changing the way the vacancy distribution increases with depth. The present vacancy distribution saturates to a plateau as seen in Figure 153. 51 45.6% C, and 8.1% Co (with a trace of TaC-NbC), or a W-to-Co ratio of 5.63-to-1. To produce a 7.35-to—l ratio within the chip, we hypothesize that for every 100 moles of tool decomposed, Aw moles of W, Ac moles of C, and AC0 moles of Co are lost due to non-dissolution mechanisms; that is, the apparent W-Co molar ratio as seen by the chip is 4i6—Aw —8-1-Aco =7.35. (50) Aw , Ac , and AC0 are fixed by one extra piece of information from measurements: When the mole fraction of intact WC lost to non-dissolution wear is equated to the residual WC found (the chips are dissolved in hydrochloric acid), it is found to be @34un) + (3517ttg/Mo) = 0.066 [Subramanian et al., 1993, Table 1], where M0 is the molecular weight of WC. Alternatively, the total number of moles of WC lost to non- dissolution wear out of 100 moles is Aw + AC = ZAW = 6.6 moles. (51) The solution to Equations (50) and (51) is that Aw = 3.3 moles and AC0 = 2.345 moles. Hence, out of every 100 moles of tool material decomposed, about 2Aw + AC0 z 9 moles are unavailable to the dissolution mechanism for the 4-m/s experiment. A similar calculation is carried out for the case of 2.5-m/s cutting speed, with ‘7.35’ in Equation (50) replaced by ‘5.18’ (Figure 25b), and with ‘6.6’ in Equation (51) replaced by ’33.6’ [Subramanian et al., 1993, Table 1]. The corresponding solution is that, out of every 100 moles of tool material decomposed, 16.8 moles of W, 16.8 moles of C, and 2.54 moles of Co are lost to non-dissolution wear at a cutting speed of 2.5 m/s. More sophisticated 52 models of non-dissolution wear can be used if information is known about the amount of tungsten lost beyond the amount that is lost as intact WC grains. 3.5.5. The rate of tool wear as a field quantity An analogous relation to Equation (42) can be written to describe the flux of material leaving the tool interface 3X- " [Di 3 l J = [decompkdecomp (X i, eq — X i ), (52) "‘001 tool where ntoo] is the outward normal distance to the tool interface, ldecomp is used to render a volumetric rate into an area flux, and kdecomp is the rate constant for the decomposition of tool material defined by (53) Agdecomp ] Rgtool kdecomp = BOVtool exp[- Figure 26a shows a plot of [decompkdecomp for ldecomp z 2.5X10_8m , Bo = 1, Vtool z 1x1013s'1 for the case of 4-m/s cutting speed, and Figure 26b shows the corresponding boundary temperature: 53 Value of surface rate constant Tool-interface temperature, of decomposition for 4-rn/s 4 mls cutting speed 10‘ cutting speed 2.5x . recs 2 A1400» 81200 17; 1.5 g E ‘97 1000 V 1 8 g 800 “o 0.5 1 1.5 “”0 0:5 1 1.5 distance from cutting edge at 10‘3 distance from cutting edge x 106 along tool-chip interface (m) along tool-chip interface (at) Figures 26a, b. The surface rate constant for tool decomposition and tool- interface temperature The similarity in functional forms suggests that the contribution to the wear contour is a direct consequence of the interfacial temperature profile, which is recognized early on by Chao and Trigger [1955, p. 1117]. Recently, Molinari and Nouari [2002] have also reported calculations showing that experimental wear profiles on the tool are consistent with the diffusion mechanism. Instead of using a boundary flux condition as reflected by Equation (42) or Equation (52), Molinari and Nouari use Dirichlet—type conditions. While the model of Molinari and Nouari is sufficient to describe the tool-interface contour, the use of a single diffusion equation for the chip region with a Dirichlet boundary condition leads to exponentially decaying profiles of tool components into the chip; this stands in contrast with the findings of Subramanian et al. The question of whether Dirichlet or flux conditions are the proper ones to be applied at the interface may well be decided by details at the interface. However, it is noted that the boundary condition of Equation (42) behaves as a 54 Dirichlet condition when the dissolution rate constant is high enough to guarantee that the interstitial solubility is reached along the interface (e.g., Section 3.5.3). 3.6. Measurement methods for checking the distribution of vacancies Macroscopic measurement methods of vacancies such as dilatometry [Paris et al., 1975] measure bulk changes in volume and are not suitable for charting spatial distributions. Evidently, highly sensitive methods operating at the level of microstructures are required. Two such methods for measuring the distribution of vacancies are field ion microscopy and the tracking of positron annihilation (which includes positron lifetime spectroscopy and gamma-ray angular correlation). Although experimental work involving these two methods has yet been applied to the machined chip, a brief description is included here for completeness. In field ion microscopy, the specimen is sharpened to a submicron-size point and a high voltage is applied to create an electric field at the tip that alternately polarizes and attracts, then ionizes and repels, the atoms of an imaging gas (e.g., neon). The gas molecules are energized according to the protrusions and indentations of the microstructure at the specimen tip, and are propelled in a radial direction towards a phosphor screen to produce an image. The fluorescent image thus created is then a magnified picture (approximately x106) of the intersection of the specimen microstructure with the surface of the needle point. Useful information is extracted through careful comparisons with crystallographic models. Lattice defects such as vacancies (and in general, point, line, and planar defects) are discernible through changes in image brightness and contrast. A mass spectrometer can also be used to identity 55 impurity atoms. Although powerful, the use of field ion microscopy requires many control experiments to eliminate visual artifacts; a general discussion and a list of references are given in [Miller and Smith, 1989; Miller et al., 1996]. In the positron method, a stream of positrons created from a radioactive source is directed into a solid. Within the bulk of the solid, these positrons are trapped by lattice imperfections. As positrons are annihilated upon contact with electrons, gamma radiation is released, and a map of the traps, and hence the lattice imperfections, is created. Because positrons are point particles, the resolution of this non-intrusive method is extremely high. However, as positrons are not trapped by interstitials, this method is limited to detection of “open-volume defects” such as vacancies [Hautojarvi, 1979; Krause-Rehberg and Leipner, 1999]. 3.7. Non-uniqueness of Frank-Tumbull representation The most important of the parameters for curve fit are a, aged,“ , and the diffusivities (Sections 3.5.1, 3.5.2, 3.5.3). Of these, Agacfivn is the parameter with the most amount of variability. Referring again to Figures 23, when Afg‘mm is less than about 35000 cal/mole, the entire contact interface from 0 to 1.8 mm achieves saturation (i.e., X i = X i.eq uniformly). As Maw,“ increases, X i becomes less than X i.eq beginning from the cutting edge and moves downstream, but curve fit at the x = 1.8-mm section can still be performed. As mentioned earlier, Mam," can be increased to levels where X i becomes uniformly less than X i, eq along the contact interface, until it becomes difficult to match the peaks of the concentration profiles of Subramanian et al. 56 To better estimate the value of Agacfivn , experimental concentration profiles are needed at different distances away from the cutting edge. The previous discussion is based on the idea that one single profile can be approached spatially in an infinite number of ways. The calibration constant A0 in Table 4 is also non-unique by at least an order of magnitude, mainly because the rate constant is dominated by the exponential terms. Further non—uniqueness of representation may arise from the non-linear nature of the governing equations. However, restrictions regarding the size of a(Section 3.5.2) and the values of the diffusivities (Section 3.5.1) are expected to hold in all circumstances. 3.8. Notes on the numerical method The finite element source code is modified from 2D programs for the heat equation taken from the text of Kwon and Bang [2000, Examples 5.9.1 and 6.6.2]. In the Oth-order problem, the main additions are the advection field and coupling between interfacial tool and chip temperatures. In the lst-order problem, the program is enlarged to model 3 scalar fields that mutually interact through advection, diffusion, and reaction. In the transient problem of relaxation, X 5 cannot be explicitly expressed as an algebraic function of X i and X V ; hence, it cannot be eliminated as in the steady-state problem. The Newton-Raphson method (Crank-Nicolson time integration with nonlinearity iteration on the solution at the new time step [Zienkiewicz and Taylor, 2000a, pp. 6-8]) proves stable not only for the differing magnitudes of coefficient matrices, but also for handling sharply varying gradients at small regions of the mesh 57 where wiggles in the solution may be present. Small wiggles are tolerated as a matter of practicality due to the use of the coarsest meshes possible for the quickest solution speeds, but with a compromise in accuracy. In separate tests, wiggles are minimized by local smoothing of any sudden changes in the way kdiss varies near the cutting edge, but the impurity profiles and distributions do not show significant changes. Because the region of interest in the mass-transfer problem measures roughly 0.2 microns by 1800 microns, elements severely elongated in the flow direction (Jr-direction) have been used. Computationally, the effect is equivalent to neglecting the x-diffusion term, which implies that the relaxation results are only valid for time scales much shorter than that which is necessary for mass-transfer in the flow-direction x. This is deemed acceptable for the geometry considered. Also, in the present case where most boundary conditions are of the flux type, the desired level of mesh refinement to minimize or avoid wiggles can be obtained by trial and error (as opposed to the case of Dirichlet boundary conditions, where limits are placed on the element Peclet number [Gresho and Lee, 1979]). 58 Chapter 4 CONCLUSIONS AND SUMMARY In ascribing to the experimental concentration profiles a physical interpretation using the Frank-Tumbull mechanism, we have made use of many assumptions, varied a number of parameters to achieve curve-fit, and calculated quantities that reflect the implications of the model. Here we provide a summary. 4.1. A list of assumptions made, with comments (1) The work/chip material is assumed incompressible. Implication: The energy equation does not involve a change in volume term, mainly for simplicity. (2) Balances of linear momentum and mass are satisfied using the assumptions of classical machining theory. In particular, uniform work and chip flow are assumed sufficient for the present emphasis on high-temperature phenomena. (3) Bridgman’s generalized entropy is assumed for a perfectly plastic work piece. Implication 1: Accessibility of states via reversible processes is not required. Implication 2: The stress power T : D is completely converted into expelled heat during plastic deformation, which is used to make sense of the meaning of absolute temperature appearing in the energy equation. 59 Implication 3: The thermodynamic state functions do not depend on kinematic parameters of plasticity, as T : D is treated as a heat source. (4) Weak coupling is assumed between thermal fields and mass-transfer fields, which allows separate consideration of the equations of energy and mass-transfer. Note: This assumption is justified by the difference in the sizes of thermal and mass diffusivities, but its applicability is limited to regimes where the tool has not worn so much that the background temperature is altered. Implication: Insular irreversibility [Bridgman, 1950] is limited to the mechanical aspect of the problem governing the background temperature field. A steady temperature field governs the values of constituent solubilities within the chip in the mass-transfer problem. (5) Chemical equilibrium is invoked in the description of mass transfer—namely, in the phrasing of the boundary flux in terms of the probabilities of tool decomposition and chip-boundary activation. Note: Our results are consistent with the work of Kramer [1979] in the following sense. The Arrhenius relation enters the flux boundary conditions as probabilities altering the value of the rate constant to the interfacial reaction. In turn, the rate constant governs the speed of rise (spatially and temporally) of the interstitial distribution towards its saturation value. However, the saturation value itself, namely the solubility, is unaffected. As a measure of the maximum amount of dissolution wear possible under all machining conditions, the solubility remains the most important factor. 60 (6) The order in which the following events occur is immaterial: (i) the system achieving a steady state in global mass transfer, and (ii) the Frank-Tumbull reaction reaching local saturation. Implication: 30/6: terms and F3 are set to 0 for the time-independent equations. (7) The concentration profiles of Subramanian et al. are calibrated in our simulations to correspond to a cut section at the end of the contact interface, which is about 1800 microns from the cutting edge. This is the location where the concentration profiles are close to achieving their maximum levels. However, simulation results can be recalibrated to match data at other sections along the tool-chip interface. (8) The concentrations of defects and impurities are considered small compared to the chip-material concentration, and the total concentration c is assumed constant. Implication: c is eliminated in the mass transfer equations in favor of mole fraction X, and data from binary phase diagrams can be used directly. (9) Tracer diffusion is assumed for W and Co within the Fe lattice. Note: This assumption is made in light of a scarcity of data with ternary and multi-component mixtures [Shewmon, 1989, Section 4.6], but it is fortunately justified by the low concentration levels actually observed. Implications: diffusivities are independent of impurity concentrations; equilibrium mole fractions of impurities are directly proportional to Arrhenius exponential terms [Shewmon, 1989, Section 2.4]; forcing terms in the species- transfer equations are simplified (Section 3.1). 61 (10) A vacancy distribution is assumed to rise from a very low value near the interface and jump quickly to a saturation value, mirroring the first rise of the concentration profile of Subramanian et al. Implication: The assumed vacancy distribution is essential in bringing about the low entry value of X s before the first rise of the humped distribution, and it reflects the statement that dissolution is “vacancy-controlled for small depth, but interstitial-controlled for large depth.” (1 l) The C0 binder atoms enter the chip region such that the original W-Co molar ratio within the tool is approximately preserved: variations are assumed to be incurred by material loss at the interface. Justification: That the W-Co molar ratio is approximately preserved is checked directly using experimental profiles (Section 3.4). Effect on simulations: The solubility of Co is related to the solubility of W by XCo,eq = ¢XW,eq (¢<1)- Implication: Knowing that the cobalt content is restricted by the availability of tungsten within the chip allows the amount of cobalt lost to mechanisms other than dissolution to be estimated. (12) The W concentration profile is assumed to decay asymptotically to zero and forces the Co concentration profile to vanish along with it (by Assumption 11). Note: This is simply a modeling decision that frees us from having to impose special assumptions to generate a buildup of X s at large depths into the chip. 62 (13) Pressure and damage are two independent ideas that have potential to explain the drastic reduction in the mass diffusivities Di and DV from table look-up values. Note: The drastic reduction in diffusivity values is necessary to maintain a rapidly-rising vacancy distribution and rapidly-decayin g interstitial distribution; in turn, the product of these distributions gives rise to the humped X 5 distribution that is matched with the experimental data of Subramanian. It is noted that the mass diffusivities used for the two cutting speeds do not differ substantially (Table 4), and are not changed for the post-machining relaxation simulations. This is viewed to be consistent with damage. (14) Because W and Co have similar atomic radii to Fe, they are assumed to exist more favorably as substitutional impurities rather than interstitial impurities. Hence, X 8, eq for W (or C0) is approximately equal to but is less than X W, eq (or X Co, eq ), whereas X i, eq for W (or C0) is a small remainder, say aX W, eq (or aX Co, eq ). ahappens to be a parameter of curve-fit that directly controls the initial condition of the post-machining relaxation (Figure 20 and its discussion). (15) The effect of pressure and vacancy concentration on the solubility of interstitials and substitutionals is not expected to change the present results drastically. Reasoning: X S is source-limited by the boundary supply of interstitials, or X 1, eq . Therefore, X s is of the same order as the maximum of X i, eq , implying that X S « X 5’ eq . What is being assumed here is that in spite of changes to X 3, eq due to pressure and vacancy concentration, X S is still 63 expected to be small in comparison to its ceiling value of X 5, eq . The parameter of importance is judged to be the curve-fit value of X i, eq at the boundary (viz., the value of a). Error is incurred if changes in X S, eq due to pressure and vacancy concentration are so drastic that the X 5 distribution is significantly changed by the changes in the solubility limit. Implication: X 8, eq is based on the solubility of tungsten in iron, which is obtained directly from a W-Fe binary phase diagram without regard to the prevalent pressure and vacancy concentration. (16) X V, eq is treated as a variable parameter of fitting with experimental data in the absence of an explicit molecular model explaining the formation of vacancies. (17) The slow decay of X S is hypothesized to be an effect of relaxation after machining. Note: Given the constraint of a small value of X i, cq , the matching the steady- state X 8 distribution to the experimental data has proved difficult. The hypothesis is then developed from the realization that the measured impurity profile is the sum of the steady-state X 5 distribution and the residual X i distribution. Computationally, relaxation gives us one more chance to improve the curve-fit. Physically, a post-machining relaxation actually occurs in the short time between the end of cutting and the quenching of the chips, although the exact time elapsed until quenching has not been reported in the experiments of Subramanian et a] [1993]. (18) Interfacial temperature jumps are controlled by the heat-transfer coefficient field, as discussed and justified in Chapter 2 and the Appendix. 4.2. A list of key parameter variations used to achieve match with experimental data (1) X i, cq (the maximum boundary value of impurity concentration) is chosen so that the residual X i distribution provides a desired alteration of the X s decay during the relaxation after the steady-state has stopped. This is controlled by the fraction 0' in Table 4. (2) X W, eq and X C0, eq are related by a fraction ¢ (Section 3.5.4). (3) Agactivn mainly controls the variation of X i along the tool-chip interface, in other words, how much too] components are admitted into the chip lattice. In turn, this affects how X S varies along the flow direction. (4) Di controls the spread of interstitials, and hence the location of the maximum of the X 8 distribution; DV is set so that given a cutting speed, the vacancy distribution applied within the primary shear zone is advected relatively unchanged across the length of the contact interface. These values of diffusivities—much lower than table lookup values—help to achieve the rise and fall of the X 5 distribution, the location Of its maximum, and its spatial spread. 65 (5) The saturation value of X V, eq affects the X 5 distribution only insofar as it affects the manner in which the X V, eq function rises, but it does not affect the tail of the X 5 distribution. (6) For post-machining relaxation, the value of kfwd determines the rate at which the Frank-Tumbull reaction occurs in relation to the rate of the diffusing interstitial species. A uniform reaction constant kfwd produces fairly good results; improvement is possible if the uniformity of kfwd is not required. 4.3. A summary of motivations, results, and quantities calculated (1) The functional form of the rate constant kdiss for the dissolution of tool constituents inside the chip is obtained by a model of dissolution that is consistent with Kramer’s earlier work. Undetermined parameters in the expression for kdiss are estimated via a curve-fitting process using the concentration profiles of Subramanian et al. [1993]. (1a) Dissolution is defined as the combined events of tool decomposition at the contact interface and subsequent mass transfer and exchange (advection, diffusion, reaction) into the chip. (lb) The boundary supply of tool constituents into the chip region is capped by the solubility of tool material within the chip. It is also influenced spatially by Arrhenius terms governing tool decomposition and the jump frequency of tool atoms across the contact interface. 66 (2) (3) (4) (5) (6) (1c) Low temperatures hinder dissolution. In particular, dissolution wear is decreased wherever the chip interface is cooler due to loose contact. The dissolved, concentration field within the chip at steady state is calculated using the proposed Frank-Tumbull mechanism, and curve fit is improved using a proposed, mass-transfer relaxation that occurs after cutting. (2a) The proportions of interstitial impurities and substitutional impurities among the total amount of impurities are given by a curve-fit parameter of the relaxation simulation. Table-lockup values of diffusivities for W and Co within Fe are too high to achieve a humped concentration profile at the observed space scales. Curve-fit values are used instead, and it is proposed that damage may be the cause for the drastic reductions in diffusivities. Table-lockup values of the solubility of Co in Fe are too high to achieve the boundary concentration of Co observed. To cap the value of Co solubility, we hypothesize that cobalt remains in roughly equal molar proportions to tungsten before and after cutting. With information on the amount of WC grains that remain intact after machining [Subramanian et al., 1993, Tables 1 & 2], the proportions of worn cobalt attributed to dissolution mechanisms and non-dissolution mechanisms are estimated. The functional form of the rate of wear for the tool is given, and it is shown to obey the same variations as the tool-interface temperature. 67 Appendix RELATIONSHIP OF INTERFACIAL TEMPERATURE JUMPS TO THE HEAT PARTITION FIELD USING A 1D PROBLEM Al. Introduction When one material slides against another, it is commonly assumed that the steady—state temperatures of the two bodies match at their contact interface. In this Appendix, we derive conditions under which interfacial temperature continuity does not hold using a 1D example. The simplicity of the 1D problem allows us to study the cases of an uncoupled interface, an interface obeying temperature continuity, and a convectively cooled interface. It will be shown that when an interfacial-temperature jump does occur, its magnitude can be expressed in terms of the heat partitioned to either side of the interface. Further, for the convectively cooled interface where frictional flux is generated, if the heat partitioned to the work is less than that for the case of temperature continuity, then the work temperature will be cooler than the tool temperature (Inequalities (A24)). A2. The 1D problem In Figure Al, the work and tool materials are schematically represented to lie along the x-axis, and they are in contact at the location shared by the work node w2 and the tool node tl: 68 frictional flux Qapplied t1 tool t2 A ‘ v v M work w2 Figure A1: Ideal 1-D representation of the work and tool materials A heat flux Q(in W/mz) is applied at the interface. The energy equations governing the work and tool domains are 2 d 0 .. d0 kwork 61:20“ “pworkcworkv xrk =0: (A1) 2 d 9 km] “2°01 =0. (A2) dx Here, k is the thermal conductivity, p is the density, 6 is the specific heat (per unit mass), v is the cutting speed, Bis the temperature, and x is the distance coordinate. The system is subject to two far-field boundary conditions: aworklxzxW1 = wl = 0, (A3) and 0tool|x=m = 9,, = 0, (A4) the application of which to the solution of Equations (A1) and (A2) yields ‘ ~ 1 exp[Pwol:kcwlprkV (x-xw1)]"l x _x mek : Aw p WC; v and 6tool = At[ Z ] . (A5), (A6) exp[ wol:k work Lwork J _ l 001 L work _ 69 Here, Aw and At are constants to be determined, Lwork = xwz —xw1 , and Ltoo] = xa — xtl. The complete solution (for AW and At) depends on the interfacial conditions applied. A2. 1. The uncoupled interface At the interface, the conservation of energy requires that 36 work d6 kwork + ktool [" AOL] = Q - (A7) ax x=xw2 dx x=xn In the absence of additional information on interfacial behavior, the assignment of heat partition is arbitrary. For convenience, an equal partitioning of heat to both work and tool is usually chosen, which is equivalent to the condition a 6 work kwork " ktoolETiqm‘) = 0 ° (A8) 8x x=xw2 dx Jr=xt1 The system of Equations (A1) and (A2), subjected to conditions (A3), (A4), (A7) and (A8), yields an exact solution of - - A \ exp[pworkcworkv (x "xwl) -l 9 k ___ Q kwork 1 (A9) wor . 2PworkC workv p[1)workéworkv ) ex Lw k L kwork or 1 _ gtool = —Q (th — x) (A10) 2ktool As mentioned, the heat partition function ,6, which for concreteness is defined to be the fraction of heat flowing into the work, is arbitrarily set to 0.5: 70 flwork = fl= 05. (Al 1) 191001 = 1 — fl: 0.5. (A12) A22. Temperature continuity The interfacial conditions in this case are kwork aawork + ktoolL— M) = Q 1 (A13) ax x=xw2 x=xt1 aworklx=xw2 _ 6tool|x=xfl = w2 _ atl : 0 . (A14) which, together with conditions (A3) and (A4), result in the following exact solution of the system of Equations (Al) and (A2): r Q .1 f " pworke workV 1 W ex Lwork kwork + ktool Pworke workv P 6 V Lt 1 exp[ wo]:k work Lwork J _ l 00 _ work _ j t (A15) - e 1 ex;{ pworkcworkv (x _ xw1)] _1 kwork 6 v exp( pwork work Lwork ] _ 1 i. k work 6work = 1 K - J Q th‘x gtool = f — [ exp£9workcworkv Lwork] W W [“001 k work + k tool Pworkéworkv 6 V Lt exp[ 9 wolzk work Lwork ] _ 1 001 ( _ work _ j ] (A16) Heat partition functions can be calculated from the fluxes evaluated at the interface, and they are 71 1 86 .6 = E[kwork gxmk P work 6 work V Lwork kwork P work C work V 6 v Lexp[ Pwork work Lwork J _1 . (A17) — - kwork exp( P work6 work V Lthk .. kwork P work c work V exp[ L l Lwork J “'1 P workc work V d k tool fltoolzl-fl‘: LL00] p- k work P workéS work V kwork e V exp( Pwork work L Lwork ] P work C work V k work A2.3. Convective cooling Lwork ] “l q and (A18) + ktool Ltool The convective-cooling interface is modeled with a heat-transfer coefficient h as follows. From the work side: 39mm = “(Tu ‘Tw2)+.5Q 8x x=xw2 k WOT From the tool side: 36001 ktool(__atT') = h(Tw2 _ t1) + (I‘mQ _ x=xu 72 (A19) (A20) The four conditions (A3), (A4), (A19) and (A20),13 when applied to the system of Equations (Al) and (A2), yield the following solution, which for simplicity, is given in terms of the constants Aw and At of Equations (A5) and (A6): fl-—::°°‘ +h Aw = w2 = 9“ -Q.(A21) / x ( ex({ P workC work V Lwork ] W W k [h + ktoo} h + P workéworkv ' p[ work ex J Ltool J P k5 k" M: wor Lwork J T 1 work ) ( f . ‘1 ) C V exr{ P work work I ork J kwork + h (1" :6) Pworkeworkv ' CX[{ P k5 kV wol: wor Lwork] _ 1 work ) ex{Pworkéworkv Lwork) W W kt IN A kwork [h + _go_ h + PworkcworkV ° LI 1 6 v 00 ) ex Pwoli'k work Lwork _1 K L ( work } The equality of the constants AW and At to the interface temperatures of the two regions A. = 611= ) ; 'Q .(A22) U can be seen by inspection of Equations (A5) and (A6). From Equations (A21) and (A22), it is seen that setting h = 0 corresponds to the uncoupled-interface solution (Equations (A9) and (A10) with ,6 = 0.5), while setting h = 00 corresponds to the temperature-continuity solution (Equations (A15) and (A16)). Hence, the heat transfer coefficient h can be used to describe intermediate behavior; it is presently used to quantify the temperature jump between tool and work: 13 The case of Q: 0 corresponds to flux continuity across the interface and can also be considered; however, the present boundary and interface conditions must be altered for the solution to be non-trivial. 73 At ’Aw = gtl _6w2 .. N C V ( exr{ Pwolzk wifrk Lwork] k (l-fl)' PworkeworkV' p ewor v ‘fl'fi exp[ work work Lwork] -l 00 _ \ kwork J . Q (A23) / f .. \ ex PworkcworkV Lw W ktool \ .. kwork ork 2 h +£— h‘i'PworkaorkV' e V “h 001 ) exp( Pwol:k work Lwork ] _ 1 K ( work ) ) Since the denominator of the last expression in Equation (A23) is always positive, the inequality holds as long as ( PworkcworkV ' K P workcwork V ' K 8X 9w2 -<- 6t] .1 .1 i k work k work P work 6 work V k Lwork work P orke okV wk w r Lwork work P work 5 work V Lwork P work 6 work V Lwork J-l Ltool (A24), ) ) lb ] + ktool (A24)2 Y ) Note that the right-hand side of Inequality (A24); is just the heat partition for the case of temperature continuity (Equation (Al7)). Hence, under the assumption of a convective- cooling interface, if the work absorbs less heat than it would have in the case of temperature continuity (Inequality (A24)2), then the work will be at a lower temperature than the tool at the interface (Inequality (A24);). The precise amount of the temperature 74 jump is given by Equation (A23). A similar result can be obtained if the direction of the previous inequalities is reversed; namely, am 2 (9” if ,62 {right-hand side of (A24); }. The heat partition ,8 in the case of convective cooling is viewed as a process parameter that depends on material properties and the details of contact. While ,6 cannot be solved explicitly in the present setup of convective cooling, its limiting behavior can be deduced. Suppose in the steady state, the tool and work temperatures are postulated to match at the interface if the work is stationary, even under convective cooling. In other words, in the limit where v = 0, we set Aw = “,2 = 6,1 = A , and from Equation (A23), arrive at kwork lim ,8 = LW" (A25) v—)O (1(on + ktoo/ ) . Lwork Ltool On the other hand, as v —> co, the interfacial temperature jump reduces to [1— lim )6) lim At —Aw = lim Tu —Tw2 = v_’°° -Q 20 whenever 13$ 1. (A26) 001 Note that for temperature continuity, limv_,°° ,6 =1 (from Equation (Al7)). For the interfacial temperature to drop from tool to work, the condition limv_,°° ,6 <1 is needed (for consistency with Inequalities (A24)); the plausibility of such a condition can be shown in the case of a heated object sliding across a cold surface, but will not be pursued here. 75 A.3. Summary Using the concept of the heat transfer coefficient, we have concluded that, as long as the heat partitioned to the work is less than that for the case of temperature continuity, then the work temperature will always be cooler than the tool temperature in the 1D problem. 76 BIBLIOGRAPHY Ackroyd, B., Akcan, N.S., Chhabra, P., Krishnamurthy, K., Madhavan, V., Chandrasekar, S., Compton, W.D. and Farris, TN, 2001. "Exploration of Contact Conditions in Machining," Proceedings of the Institution of Mechanical Engineers, Journal of Engineering Manufacture, Proceedings Part B, 215(B4):493-507. Bird, R.B., Stewart, W.E. and Lightfoot, EN, 2002. Transport Phenomena, 2nd edition. .1. Wiley, New York, xii, 895 pages. Bridgman, P.W., 1950. "The Thermodynamics of Plastic Deformation and Generalized Entropy," Review of Modern Physics, 22(1):56-63. Chao, B.T. and Trigger, KL, 1955. "Temperature Distribution at the Tool-Chip Interface in Metal Cutting," Transactions of the ASME, 77:1107-1121 (Discussion pp. 1 1 19-1 121). Chao, B.T. and Trigger, K.J., 1958. "Temperature Distribution at Tool-Chip and Tool- Work Interface in Metal Cutting," Transactions of the ASME, 80:311-320 (Discussion pp. 318-320). Cook, NH. and Nayak, RN, 1966. "The Thermal Mechanics of Tool Wear," Journal of Engineering for Industry, Transactions of the ASME, 88:93-100. DeVries, W.R., 1992. Analysis of Material Removal Processes. Springer Texts in Mechanical Engineering. Springer-Verlag, New York, xvii, 254 pages. Doyle, E.D., Horne, J .G. and Tabor, D., 1979. "Frictional Interactions between Chip and Rake Face in Continuous Chip Formation," Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 366(1725): 173-183. Elliott, J .F., Gleiser, M., Ramakrishna, V. and The American Iron and Steel Institute, 1963. Thermochemistry for Steelmaking, Volume II. Addison-Wesley Series in the Engineering Sciences - Metallurgy and Materials. Addison-Wesley Publishing Company, Reading, MA 77 Frank, RC. and Tumbull, D., 1956. "Mechanism of Diffusion of Copper in Germanium," Physical Review, 104(3):617-618. Giese, A., Stolwijk, NA. and Bracht, H., 2000. "Double-Hump Diffusion Profiles of Copper and Nickel in Germanium Wafers Yielding Vacancy-Related Information," Applied Physics Letters, 77(5):642-644. Gresho, P. and Lee, R.L., 1979. "Don't Suppress the Wiggles-They are Telling You Something." In: T.J.R. Hughes (Editor), Finite Element Methods for Convection Dominated Flows: Presented at the Winter Annual Meeting of the American Society of Mechanical Engineers, New York, New York, December 2-7, 1979. AMD (Applied Mechanics Division), v. 34. ASME, New York, N.Y., pp. 37-61. Hansen, M., (prepared with the cooperation of K. Anderko) and Armour Research Foundation, 1958. Constitution of Binary Alloys, 2nd edition. Metallurgy and Metallurgical Engineering Series. McGraw-Hill, New York, 1305 pages (kept up- to—date with two supplements: (1) Rodney P. Elliott and HT Research Institue, 1965. Constitution of Binary Alloys. First Supplement, McGraw-Hill Series in Materials Science and Engineering, New York, 877 pages. (2) Francis A. Shunk and HT Research Institue, 1969. Constitution of Binary Alloys. Second Supplement, McGraw-Hill Series in Materials Science and Engineering, New York, X], 720 pages. Errata for earlier volumes: pp. xxxv-xxxviii.) Hautojarvi, P. (Editor), 1979. Positrons in Solids. Topics in Current Physics, v. 12. Springer-Verlag, Berlin; New York, xi, 255 pages. Kim, W. and Kwon, P., 2001. "Understanding the Mechanisms of Crater Wear," N AMRC XXD(, 29th North American Manufacturing Research Conference, University of Florida Hotel and Conference Center, Gainsville, FL. Komanduri, R. and Hou, Z.B., 2000. "Thermal Modeling of the Metal Cutting Process. Part I - Temperature Rise Distribution due to Shear Plane Heat Source," International Journal of Mechanical Sciences, 42: 1715-1752. Komanduri, R. and Hou, Z.B., 2001a. "Thermal Modeling of the Metal Cutting Process - Part 11: Temperature Rise Distribution due to Frictional Heat Source at the Tool- Chip Interface," International Journal of Mechanical Sciences, 43:57-88. 78 Komanduri, R. and Hou, Z.B., 2001b. "Thermal Modeling of the Metal Cutting Process - Part IH: Temperature Rise Distribution due to Combined Effects of Shear Plane Heat Source and the Tool-Chip Interface Frictional Heat Source," International Journal of Mechanical Sciences, 43:89- 107. Kramer, B.M., 1979. An Analytical Approach to Tool Wear Prediction. Ph. D. Thesis, Department of Mechanical Engineering, Massachusetts Institute of Technology, 132 pages. Kramer, B.M. and Judd, P.K., 1985. "Computational Design of Wear Coatings," Journal of Vacuum Science & Technology A, Vacuum, Surfaces, and Films, 3(6):2439- 2444. Kramer, B.M. and Suh, NP, 1980. "Tool Wear by Solution: A Quantitative Understanding," Journal of Engineering for Industry, Transactions of the ASME, 102:303-309. Krause-Rehberg, R. and Leipner, HS, 1999. Positron Annihilation in Semiconductors: Defect Studies. Springer Series in Solid-State Sciences, 127. Springer, Berlin; New York, xv, 378 pages. Kwon, Y.W. and Bang, H., 2000. The Finite Element Method Using MATLAB, 2nd edition. CRC Mechanical Engineering Series. CRC Press, Boca Raton, FL, 607 pages. Mase, GT. and Mase, GE, 1999. Continuum Mechanics for Engineers, 2nd edition. CRC Press, Boca Raton, FL, 377 p. pages. Miller, M.K., Cerezo, A., Hetherington, M.G. and Smith, G.D.W., 1996. Atom Probe Field Ion Microscopy. Monographs on the Physics and Chemistry of Materials, 52. Clarendon Press; Oxford University Press, Oxford; New York, xi, 4 plates, 509 pages. Miller, MK. and Smith, G.D.W., 1989. Atom Probe Microanalysis: Principles and Applications to Materials Problems. Materials Research Society, Pittsburgh, Pa., xiv, 278 pages. Molinari, A. and Nouari, M., 2002. "Modeling of Tool Wear by Diffusion in Metal Cutting," Wear, 252(1-2): 135-149. 79 Naerheim, Y. and Trent, B.M., 1977. "Diffusion Wear of Cemented Carbide Tools when Cutting Steel at High Temperatures," Metals Technology, 4(12):548-556. Paris, D., Lesbats, P. and Levy, J ., 1975. "An Investigation of the Distribution of Vacancies in an Ordered Fe-Al Alloy by Field Ion Microscopy," Scripta Metallurgica, 9: 1373-1378. Rabinowicz, E., 1965. Friction and Wear of Materials. Wiley Series on the Science and Technology of Materials. John Wiley & Sons, Inc., New York, 244 pages. Rabinowicz, E., 1977. "Abrasive Wear Resistance as a Materials Test," Lubrication Engineering, 33(7):378-381. Rabinowicz, E., Dunn, LA. and Russell, P.G., 1961. "A Study of Abrasive Wear Under Three-Body Conditions," Wear, 4:345-355. Shaw, M.C., 1984. Metal Cutting Principles. Oxford Series on Advanced Manufacturing (No. 3). Clarendon Press; Oxford University Press, London; New York, xxii, 594 pages. Shewmon, P.G., 1963. Difiusion in Solids. McGraw-Hill Series in Materials Science and Engineering. McGraw-Hill, New York, 203 pages. Shewmon, P.G., 1989. Difi‘usion in Solids, 2nd edition. Minerals, Metals & Materials Society, Warrendale, Pa., 246 pages. Sturge, M.D., 1959. "A Note on the Theory of Diffusion of Copper in Germanium," Proceedings of the Physical Society, 73(2):297-306, (ioumal also known as the Proceedings of the Physical Society of London from 1874-1925). Subramanian, S.V., Ingle, SS. and Kay, D.A.R., 1993. "Design of Coatings to Minimize Tool Crater Wear," Surface & Coatings Technology, 61(1-3):293-299. Suh, NP, 1986. Tribophysics. Prentice-Hall, Englewood Cliffs, NJ, xiv, 498 pages. 80 Tay, A.O., Stevenson, MG. and de Vahl Davis, G., 1974. "Using the Finite Element Method to Determine Temperature Distributions in Orthogonal Machining," Proceedings - The Institution of Mechanical Engineers, 188:627-638. Trent, B.M., 1967. "Conditions of Seizure at the Chip-Tool Interface," Machinability. 151 Special Report 94. Proceedings of the Conference on Machinability jointly organized by the Iron and Steel Institute and the Institute of Metals, in association with the Institution of Mechanical Engineers, the Institution of Production Engineers, and the Institution of Metallurgists, held at the Royal Commonwealth Society, Craven Street, London WC 2, on 4-6 October 1965, pp. 11-18 (article makes reference to Fig. 13 on p. 182 of the same Proceedings [Trent, "The Relationship between Machinability and Tool Wear," ibid., pp. 179-184]; panel discussion: pp. 55-63). Wallace, P.W. and Boothroyd, G., 1964. "Tool Forces and Tool-Chip Friction in Orthogonal Machining," The Journal of Mechanical Engineering Science, 6( 1 ):74-87. Zienkiewicz, CC. and Taylor, R.L., 2000a. The Finite Element Method, Volume 2, Solid Mechanics, 5th edition. Butterworth-Heinemann, Oxford; Boston Zienkiewicz, CC. and Taylor, R.L., 2000b. The Finite Element Method, Volume 3, Fluid Dynamics, 5th edition. Butterworth-Heinemann, Oxford; Boston Zorev, N.N., 1958. "Results of Work in the Field of the Mechanics of the Metal Cutting Process," Proceedings of the Conference on Technology of Engineering Manufacture, London: 25th to 27th March, 1958. Session 3, Machining of Materials. Institution of Mechanical Engineers (Great Britain), pp. 255-266, plus Plate 4 (Figs. 54.1a-f), Plate 5 (Figs. 54.7 & 54.9a-b), Plate 6 (Fig. 54.8), & Plate 7 (Figs. 54.18a-b, 54.19, 54.20, & 54.26). Plates (not paginated) are placed between p. 406 and p. 407. Reports, Discussion, Communications, & Authors' Replies for Session 3 are contained in pp. 374-405. 81 Fl L l'riluljjjl 324 lgljyjjrlyg ht'WWmWL m"»‘a.na.. _-_- L44.- A .7 l