surface defect machining

Upload: saurav-goel

Post on 01-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/9/2019 Surface defect machining

    1/12

    Saurav Goel1School of Mechanical

    and Aerospace Engineering,Queen’s University, Belfast BT95AH, UK

    e-mail:[email protected]

    Waleed Bin RashidInstitute of Mechanical, Process

    and Energy Engineering,Heriot-Watt University,

    Edinburgh EH144AS, UK

    Xichun LuoDepartment of Design, Manufacture

    and Engineering Management,University of Strathclyde, Glasgow G11XQ, UK

    Anupam AgrawalDepartment of Business Administration,

    University of Illinoisat Urbana Champaign, IL 61820

    V. K. JainDepartment of Mechanical Engineering,

    Indian Institute of Technology,Kanpur 208016, India

    A Theoretical Assessmentof Surface Defect Machiningand Hot Machining of

    Nanocrystalline Silicon Carbide In this paper, a newly proposed machining method named “surface defect machining”(SDM) was explored for machining of nanocrystalline beta silicon carbide (3C-SiC) at 300 K using MD simulation. The results were compared with isothermal high temperaturemachining at 1200 K under the same machining parameters, emulating ductile modemicro laser assisted machining ( l -LAM) and with conventional cutting at 300K. In theSDM simulation, surface defects were generated on the top of the (010) surface of the 3C-SiC work piece prior to cutting, and the workpiece was then cut along the h100 idirection using a single point diamond cutting tool at a cutting speed of 10 m/s. Cutting forces, subsurface deformation layer depth, temperature in the shear zone, shear planeangle and friction coefcient were used to characterize the response of the workpiece.Simulation results showed that SDM provides a unique advantage of decreased shear plane angle which eases the shearing action. This in turn causes an increased value of average coefcient of friction in contrast to the isothermal cutting (carried at 1200 K)

    and normal cutting (carried at 300 K). The increase of friction coefcient, however, was found to aid the cutting action of the tool due to an intermittent dropping in the cutting forces, lowering stresses on the cutting tool and reduced operational temperature. Analy-sis shows that the introduction of surface defects prior to conventional machining can bea viable choice for machining a wide range of ceramics, hard steels and composites com- pared to hot machining. [DOI: 10.1115/1.4026297]

    Keywords: surface defect machining, MD simulation, nanometric cutting, beta siliconcarbide

    1 Introduction“Surface defect machining” (SDM) [ 1] is a recently proposed

    method of machining that aims to obtain better quality of amachined product at lower costs. This method utilizes predenedand machined surface defects on the workpiece to ease materialremoval. The central idea of this method is to generate surfacedefects in the form of a series of holes on the top surface of theworkpiece prior to the actual machining operation. The presenceof these defects reduces the strength of the workpiece which inturn aids to lower the cutting resistance during machining. Recentexperimental trials [ 1] and numerical simulations [ 2] on hardsteels have shown some very interesting and salient features of theSDM method such as lower machining forces, reduction in overalltemperature in the cutting zone, reduced machining stresses andincreased chip ow velocity. SDM machining provides a productwith better surface integrity compared to that obtained using con-ventional hard turning.

    In this work, the applicability of SDM for machining siliconcarbide is explored. Silicon carbide (SiC) is an extremely hardand brittle nonoxide ceramic material, and its properties and man-ufacturing methods are being rigorously researched [ 3 – 6]. It hasbeen demonstrated that due to its superior properties, such aschemical inertness, high thermal conductivity, high carrier satura-tion velocity, high specic stiffness ( E/q), and high-temperatureresistance, SiC is an appropriate choice to replace silicon for advanced ultra precision engineering applications especially in the

    electronic industry [ 7]. SiC is also recognized as a potential candi-date for quantum computing applications as a substitute for dia-mond [ 8], in space-based laser mirrors [ 9] and for the

    development of thermal protection system (TPS) materials for defence applications [ 10]. Demand of SiC is growing further inweapons, aerospace, microelectronic and biomedical applicationsas well as in “big-science” programmes such as the EuropeanExtremely Large Telescope (E-ELT), the Atacama Large Milli-meter/submillimeter Array (ALMA), and next generation extremeultraviolet (EUV) lithography steppers. SiC is also nding amaz-ing applications in biomedical sector especially as being a semi-conductor material because of being more biocompatible thansilicon [ 11]. Traditional orthopaedic materials such as cobaltchrome (CoCr), stainless steel and titanium, on account of beinglow wear and oxidation resistant, succumb to bone loss whichcauses implant loosening resulting in a reactive implant surface.Contrarily, SiC is capable of permanently integrating into the newbone growth on account of low wear debris and metallosis and is

    thus very effective as a coating for stents to enhance hemocompat-ibility and as a coating for prosthetic-bearing surfaces anduncemented joint prosthetics [ 12].

    In a natural state, SiC exhibits one-dimensional polymorphism:all polytypes have the same tetrahedral arrangement of Si and Catoms but different stacking sequences. It is due to this reason thatalmost 250 polytypes of silicon carbide (SiC) have been recog-nized to date [ 13]. Across all other polytypes, two major poly-morphs are a-SiC and b-SiC with hexagonal and zinc-blendelattice structures, respectively. The main engineering properties of b-SiC (3C-SiC) and a-SiC (6H-SiC and 4H-SiC) have alreadybeen summarized elsewhere [ 14]. 3C-SiC possess extremely highhardness and low fracture toughness and is therefore very difcultto manufacture [ 15]. Considering the high hardness of 3C-SiC, the

    1Corresponding author.Contributed by the Manufacturing Engineering of ASME for publication in the

    JOURNAL OF M ANUFACTURING SCIENCE AND ENGINEERING . Manuscript received August7, 2012; nal manuscript received December 18, 2013; published online February10, 2014. Assoc. Editor: Suhas Joshi.

    Journal of Manufacturing Science and Engineering APRIL 2014, Vol. 136 / 021015-1Copyright VC 2014 by ASME

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    2/12

  • 8/9/2019 Surface defect machining

    3/12

  • 8/9/2019 Surface defect machining

    4/12

    conditions along the z direction. A snapshot from the MD

    simulation after equilibration process (with and without sur-face defects in the 3C-SiC workpiece ) is shown in Figs. 5and 6 where the red and grey colors 3 correspond to siliconand carbon atoms in the workpiece and yellow color repre-sents carbon atoms within the diamond cutting tool,respectively.

    An important constraint of MD simulation studies is that theyare computationally expensive, and therefore use of high cuttingspeeds is frequent in MD simulation studies, e.g., 500–2500m/scutting speed was used by Belak et al. [ 57,58] and Komanduriet al. [ 59,60], 150–400m/s was used by Wang et al. [ 61] andLiang et al. [ 62] and, 70–100m/s was used by Noreyan et al.[63,64], Rentsch et al. [ 65] and Goel et al. [ 4,15,53]. Although,these investigations have been successful to capture key insightsof the cutting process but in the current investigation, high cutting

    speed could have affected the sensitivity of the results, particu-larly when cutting of the same conguration was to be comparedat 300 K and 1200 K. Therefore, current simulations were per-formed at a more realistic cutting speed of 10 m/s. This wasaccomplished using parallel computing through MPI interface.The calculation time for each simulation case depends on themodel size, cutting speed, cutting distance, and the number of CPUs used.

    3 Results and DiscussionsThis section covers observations and discussion of the signi-

    cance of the MD simulation results: cutting forces, chip morphol-ogy, stresses and temperature in the cutting zone.

    Fig. 3 Stresses in the cutting zone [ 5]

    Fig. 4 Potential energy function for molecular interactions inthe molecular mechanics approximation [ 54]

    3Readers are requested to refer to the web based version of this article for correctinterpretation of the colour legends.

    021015-4 / Vol. 136, APRIL 2014 Transactions of the ASME

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    5/12

    3.1 Cutting Forces and Friction Coefficient. Figure 7shows schematically the major parameters which tend to inuencethe outcome of a machining process especially during nanometriccutting of anisotropic materials such as 3C-SiC [ 50]. As shown inthe respective 2D model, two coplanar cutting forces acting on thecutting tool fundamentally govern the cutting action of the tool,i.e., tangential cutting force ( Fc) and thrust force ( Ft ). Thrust forcepredominantly inuences surface error as it tends to separate thetool away from the workpiece, while tangential force causes dis-placements in the cut chip thickness thereby relating to chatter. Acomparison of the evolution of both these forces ( Fc and Ft )obtained from the MD simulation is compared in Figs. 8 and 9 for all the three cases studied: (i) Nanometric cutting at 300 K, (ii)

    nanometric cutting at 1200 K, and (iii) cutting with surface defectmachining at 300 K.It can be seen from Figs. 8 and 9 that the magnitudes of both

    forces are signicantly higher in the conventional nanometric

    Table 2 Process variables used for performing the MD simulation

    Workpiece material Number of atoms in the workpieceNumber of atoms

    in the diamond cutting tool

    3C-SiC without holes(14.26 nm 4.6345 nm 4.278 nm)

    28,170 21,192

    3C-SiC with surface defects (holes)14.26 nm 4.6345 nm 4.278 nm)

    27,782 21,192

    Equilibrium lattice parameters for 3C-SiC: a ¼ 4.36 Å ;

    a ¼ b ¼ c ¼ 90 degDetails of surface defects (holes):Total number 7Diameter of each hole 0.713 nmDepth of each hole 1.426 nmCrystal orientation of the workpiece (010)Crystal orientation of diamond tool CubicCutting direction h100iCutting edge radius (nm) 2.297Uncut chip thickness/in-feed (nm) 1.3126Cutting tool rake and clearance angle 25 deg and 10 degEquilibration temperature 300 KelvinHot machining temperature 1200 KelvinCutting velocity 10 m/sTimestep 0.5 fs

    Fig. 5 Snapshot from MD simulation for 3C-SiC specimenwithout creation of surface defects

    Fig. 6 Snapshot from MD simulation for 3C-SiC specimen with surface defects ontop

    Journal of Manufacturing Science and Engineering APRIL 2014, Vol. 136 / 021015-5

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    6/12

    cutting performed at 300 K. However, during nanometric cuttingat elevated temperature of 1200 K, both tangential cutting forcesand thrust forces reduce, albeit, to a lesser extent. On the other hand, a noticeable reduction, especially in thrust forces, can beseen during nanometric cutting in the case of SDM. It can be

    noted here that the extent of reduction in cutting forces duringSDM will depend on various parameters such as the number of holes, dimension of holes, interspacing of holes, shape of holes,etc. However, since the reduction in the cutting forces is of inter-mittent nature, the cutting forces and thus stresses on the cuttingtool would be relieved as soon as the cutting tool met a hole (sur-face defect). Table 3 summarizes various results obtained fromthe simulation under different machining conditions, i.e., averagecutting forces, friction coefcient and resultant cutting forces.

    The data in Table 3 show that the resultant cutting force reducesfrom a value of 1449.64 nN to 1402.13 nN when the machiningwas done at 1200 K instead of at 300 K, signifying a reduction inthe cutting resistance of single crystal 3C-SiC by 3.27% at1200 K. However, the extent of this reduction is higher during the

    surface defect machining process as the resultant forces drops to1281.19 nN, i.e., a signicant reduction of 11.62% compared tonormal nanometric cutting at 300 K. It is very interesting to notehere that while the cutting forces reduced, a similar trend is notvisible in the coefcient of friction. Compared to the nanometric

    cutting results at 300 K, the coefcient of friction reduced by8.11% when the cutting was performed at 1200K. On the con-trary, surface defect machining causes an increase in the frictioncoefcient by 2.68%. This suggests that a different mechanism of chip formation is associated with the proposed SDM method. Thisphenomenon is discussed in detail in Sec. 3.2.

    A comprehensive experimental work on surface defect machin-ing on single crystal 3C-SiC is still underway, primarily due to thefact that a sufciently larger size specimen of single crystal 3C-SiC is not available till date. However, experimental trials usingSDM method have already been done on AISI 4340 steel (hardenedupto 69 HRC) with diameter and depth of holes being 0.9mm and0.1 mm, respectively [ 1]. The sizes of the holes in the experimentalinvestigation were bigger than the size used in the MD simulation

    Fig. 7 Schematic diagram of chip formation during single point diamond turning[50 ]

    Fig. 8 Tangential cutting forces during nanometric cutting of3C-SiC in three cases

    Fig. 9 Thrust forces during nanometric cutting of 3C-SiC inthree cases

    021015-6 / Vol. 136, APRIL 2014 Transactions of the ASME

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    7/12

    (MD cannot reach millimetre length scales due to restrictive com-

    putational speeds). However, the intent here is not to compare MDsimulation with experimental observations but to understand theprocess and features of the cutting mechanism. During the experi-mental investigation, an improved average surface roughness (Ra)value of 0.227 l m from SDM machining was obtained in compari-son to 0.452 l m from conventional machining using a CBN tool[1]. A more important outcome from the experiments was an inter-mittent reduction in the cutting forces as shown in Fig. 10.

    As evident from Fig. 10, the cutting tool experiences intermit-tent relaxation in the cutting load during the cutting process whensurface defects are met by the cutting tool. This causes a steepreduction in the cutting forces. This intermittent reduction in thecutting load is favorable for tool longevity as it aids in the reduc-tion of the local temperature at the cutting edge (we expand on

    this later). This trend in the variation of the cutting load is more

    reminiscent of vibration assisted machining and in this sense aqualitative comparison between these two processes can be drawnwhich is shown in Table 4.

    3.2 Chip Morphology. Figure 11 shows a superimposedimage and a comparison of the chip morphology in all the threecases investigated. On comparing nanometric cutting at 1200 Kwith conventional nanometric cutting at 300 K, it can be seen thatthe curliness of the chip has seemingly increased which is plausi-ble due to the increased plasticity of SiC at high temperature of 1200 K. However, the shear plane angle appears unchanged,unlike SDM. In the case of SDM, the cut chip thickness has

    Table 3 Comparison of cutting forces and friction coefcient

    S.N. Machining conditionAverage tangentialcutting forces ( Fc)

    Average thrustforces ( Ft )

    Average resultantforces ( (F t

    2 þ Fc2))Average friction

    coefficient ( Fc/F t )

    1 Normal machining (300 K) 835 nN 1185 nN 1449.64 nN 0.70462 Hot machining (1200 K) 762 nN 1177 nN 1402.13 nN 0.64743 Surface defect machining (300 K) 751 nN 1038 nN 1281.19 nN 0.7235

    Fig. 10 Cutting forces using ( a ) normal hard turning and ( b ) surface defect machining [ 1]

    Table 4 Comparison between surface defect machining (SDM) and vibration assisted machining (VAM)

    Vibration assisted machining (VAM) andsurface defects machining (SDM) Similarities Differences

    Cutting forces on tool Reduced cutting forces provide better surface nish and tool longevity.

    Not applicable

    Overall cutting load on tool Not applicable In VAM, periodic reduction in cutting load occurs at speciedamplitude, whereas in SDM cutting load reduces wheredislocations in the form of holes are encountered.

    Volume of material removal Not applicable Although, tool is periodically rotated to reduce the cutting load, thetotal material to be removed during VAM process remains

    unchanged. In SDM, due to the vacancies made in the form of holes, some of the volume of the material to be removed reduces.

    Tool contact with chips Not applicable In VAM cutting tool loses contact with the chips on specied am-plitude, whereas in SDM cutting chips remains in continuous con-tact with the tool.

    Operational time Not applicable No cutting action took place while the tool is disengaged in VAM,whereas in SDM continuous cutting takes place.

    Requirement of machine tool Not applicable Separate machine tool required to execute VAM whereas with anaddition of independent process, conventional machine tool isgood enough for SDM process.

    Journal of Manufacturing Science and Engineering APRIL 2014, Vol. 136 / 021015-7

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    8/12

    increased and thus the shear plane angle has decreased. This wasveried using the following equation

    tan / ¼ r cos a1 r sin a

    (4)

    where u is shear plane angle, a is tool rake angle, and r is the ratioof uncut chip thickness and cut chip thickness.

    Table 5 shows a decrease in shear plane angle from a value of 21.28deg to 20.66 deg using SDM process compared to nanomet-ric cutting at 300 K. A decrease in the value of shear plane angleunder the same machining parameters shows the dominance of tangential cutting forces over thrust forces justifying the enhancedcutting action of the tool. This corroborates to the increased forceratio as seen earlier in Table 3 during the case of SDM suggestingthe dominance of tangential cutting force to be the reason of theincrease in friction coefcient (which improves the cuttingaction).

    Figure 12 shows the measurement of the cut chip thickness andhighlights the variation in the subsurface crystal deformation lat-tice layer depth. It is interesting to note that the subsurface crystaldeformation lattice layer depth becomes wider while cutting at300K, while it becomes a little deeper while cutting at 1200 K.

    Moreover, the extent of the deformation of crystal layer under-neath the nished surface is more pronounced in both these casescompared to that in the SDM operation. It can be postulated thathigh temperature weakens the bonding forces between the atomsand hence, the atoms could easily be deformed without havingmuch inuence on the neighbor atoms. Therefore, the deformationdid not become wider and remained concentrated under thewake of the tool under the inuence of high deviatoric stresses.Contrarily, SDM process shows minimal subsurface deformation.The waviness of the nished surface also seems to have decreasedduring SDM. The defects generated for the purpose of SDM sig-nicantly weaken the material, which in turn reduces the bondingstrength of the atoms in the area of uncut chip thickness withoutdisturbing the subsurface. Also, a discontinuity in the material and

    the consequent lack of resistance to the deformation of the atomsby the adjacent atoms makes the shearing process more preferen-tial. Eventually, the material removal becomes easier. This iscompounded by the fact that the cutting tool is relieved from high

    Table 5 Comparison of chip morphology and shear angleunder different machining conditions

    S.N.Machiningcondition

    Ratio of uncut chipthickness to cut chip

    thickness (r)Shear plane

    angle ( u )

    1 300 K 0.525 21.28 deg2 1200 K 0.525 21.28 deg3 SDM at 300 K 0.505 20.66 deg

    Fig. 11 Chip morphology of 3C-SiC while cutting the workpiece after tool advan-ces to 8.3nm

    Fig. 12 Subsurface crystal lattice deformation of 3C-SiC aftertool advances to 8.3 nm

    021015-8 / Vol. 136, APRIL 2014 Transactions of the ASME

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    9/12

  • 8/9/2019 Surface defect machining

    10/12

    of a layer of quasi-continuous material. Particles of this materialtrapped below the tool continue to rub the machined surface anddestroy the remains of the feed marks left by the cutting tool.Therefore, it appears that in addition to the SDM machiningmechanism occurring at the rake face, a simultaneous polishingmechanism also proceeds with the travel of the cutting tool.

    4 ConclusionsA new machining approach named surface defect machining

    (SDM) for machining single crystal 3C-SiC at nanoscale has beentested using MD simulation and compared with ductile modemicro laser assisted machining at 1200K and conventionalmachining at 300K. The motivation of this newly developedmethod was the anticipation that the surface defects generated onthe workpiece can allow easy shearing of the material. The com-prehensive results obtained from the simulation can be concludedas follows:

    (1) The presence of premachined surface defects improves themachinability of difcult-to-machine materials through areduction in shear plane angle and shear plane area thus per-mitting reduced side ow with less metallurgical transforma-tions on the nished machined surface and the subsurface.

    (2) Surface defects cause a reduction in the average cuttingforce which relieves the cutting tool of the cutting load

    intermittently. This was attributed to the lowered bondingstrength between the workpiece atoms. While this processis reminiscent of a vibration assisted machining process,the added advantage is that it does not require any separateattachment.

    (3) The extent of subsurface crystal deformation lattice layer depth was found to be minimal in the case of surface defectmachining followed by high temperature cutting whereascutting performed at 300 K showed maximum depth of thesubsurface deformation. A relatively higher temperature onthe tool cutting edge was found as an inherent characteristic

    with the l -LAM process in contrast to SDM process. Con-sequently, l -LAM process presents the risk of acceleratedgraphitization of the diamond tools. This is because hightemperature nanometric cutting causes transfer of heat fromthe bulk of the workpiece to the cutting tool which maycompromise the life of the diamond tools. On the contrary,SDM helps to reduce the temperature at the tool cuttingedge compared to the high temperature cutting and the nor-mal nanometric cutting at 300 K.

    AcknowledgmentAuthors acknowledge the funding support of Ministry of Higher

    Education, Kingdom of Saudi Arabia for funding the Ph.D of

    Fig. 16 MD simulation showing various stages of each machining action [ 2]

    021015-10 / Vol. 136, APRIL 2014 Transactions of the ASME

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

  • 8/9/2019 Surface defect machining

    11/12

    W.B.R. and an additional funding from J. M. Lessells travelscholarship from the Royal Society of Edinburgh (2013 RSE/J. M.Lessells Travel Scholarship), International Research Fellowshipaccount of Queen’s University, Belfast and an EPSRC researchgrant (Ref: EP/K018345/1).

    References[1] Rashid, W. B., Goel, S., Luo, X., and Ritchie, J. M., 2013, “An Experimental

    Investigation for the Improvement of Attainable Surface RoughnessDuring Hard Turning Process,” Proc. Inst. Mech. Eng., Part B , 227(2), pp.338–342.

    [2] Rashid, W. B., Goel, S., Luo, X., and Ritchie, J. M., 2013, “The Developmentof a Surface Defect Machining Method for Hard Turning Processes,” Wear ,302(1–2), pp. 1124–1135.

    [3] Goel, S., Stukowski, A., Luo, X., Agrawal, A., and Reuben, R. L., 2013,“Anisotropy of Single-Crystal 3C–SiC During Nanometric Cutting,” Model.Simul. Mater. Sci. Eng. , 21(6), p. 065004.

    [4] Goel, S., Luo, X., Reuben, R. L., Rashid, W. B., and Sun, J., 2012, “ SinglePoint Diamond Turning of Single Crystal Silicon Carbide: Molecular DynamicSimulation Study ,” Key Engineering Materials, Liverpool. pp. 150–155.

    [5] Goel, S., Luo, X., and Reuben, R. L., 2012, “Shear Instability of Nanocrystal-line Silicon Carbide During Nanometric Cutting,” Appl. Phys. Lett. , 100 (23),p. 231902.

    [6] Goel, S., Luo, X., Comley, P., Reuben, R. L., and Cox, A., 2013,“Brittle–Ductile Transition During Diamond Turning of Single Crystal SiliconCarbide,” Int. J. Mach. Tools Manuf. , 65, pp. 15–21.

    [7] Neudeck, P. G., 2000, “SiC Technology,” The VLSI Handbook , B. Raton, ed.,CRC Press , Boca Raton, FL, pp. 6.1–6.24.

    [8] Dzurak, A., 2011, “Quantum Computing: Diamond and Silicon Converge,”Nature , 479(7371), pp. 47–48.

    [9] Shore, P., Cunningham, C., DeBra, D., Evans, C., Hough, J., Gilmozzi, R.,Kunzmann, H., Morantz, P., and Tonnellier, X., 2010, “Precision Engineeringfor Astronomy and Gravity Science,” CIRP Ann. , 59(2), pp. 694–716.

    [10] Newsome, D. A., Sengupta, D., Foroutan, H., Russo, M. F., and van Duin, A.C., 2012, Oxidation of Silicon Carbide by O 2 and H 2O: A ReaxFF ReactiveMolecular Dynamics Study, Part I,” J . Phys. Chem. C , 116(30), pp.16111–16121.

    [11] Coletti, C., Jaroszeski, M., Pallaoro, A., Hoff, A., Iannotta, S., and Saddow, S.,2007, “Biocompatibility and Wettability of Crystalline SiC and Si Surfaces,”29th Annual International Conference of the IEEE. Engineering in Medicineand Biology Society , EMBS IEEE.

    [12] Goel, S., Yan, J., Luo, X., and Agrawal, A., 2013, “Incipient Plasticity in 4H-SiC During Quasistatic Nanoindentation,” J. Mech. Behav. of Biomed.l Materi-als, (in press).

    [13] Perrone, D., 2007, “Process and Characterisation Techniques on 4H—SiliconCarbide,” Micronanotechnology, Ph.D. thesis, Politecnico di Torino, Torino,Italy.

    [14] Luo, X., Goel, S., and Reuben, R. L., 2012, “A Quantitative Assessment of Nanometric Machinability of Major Polytypes of Single Crystal SiliconCarbide,” J. Europ. Ceram. Soc. , 32(12), pp. 3423–3434.

    [15] Goel, S., Luo, X., and Reuben, R. L., 2012, “Molecular Dynamics SimulationModel for the Quantitative Assessment of Tool Wear During Single Point Dia-mond Turning of Cubic Silicon Carbide,” Comput. Mater. Sci. , 51(1), pp.402–408.

    [16] Ravindra, D., 2011, “Ductile Mode Material Removal of Ceramics and Semi-conductors,” Ph.D. thesis, Western Michigan University, Kalamazoo, MI, p. 312.

    [17] Shayan, A. R., Poyraz, H. B., Ravindra, D., Ghantasala, M., and Patten, J. A.,2009, “Force Analysis, Mechanical Energy and Laser Heating Evaluation of Scratch Tests on Silicon Carbide (4H-SiC) in Micro-Laser Assisted Machining ( l -LAM) Process” ASME Conf. Proc. , MSEC2009 -84207, pp. 827–832.

    [18] Brinksmeier, E., and Gl €abe, R., 2001, “Advances in Precision Machining of Steel,” CIRP Ann. , 50(1), pp. 385–388.

    [19] Evans, C., and Bryan, J. B., 1991, “Cryogenic Diamond Turning of StainlessSteel,” CIRP Ann. , 40(1), pp. 571–575.

    [20] Casstevens, J. M., 1983, “Diamond Turning of Steel in Carbon-SaturatedAtmospheres,” Precis. Eng. , 5(1), pp. 9–15.

    [21] Brehl, D. E., and Dow, T. A., 2008, “Review of Vibration-Assisted Machining,”

    Precis. Eng. , 32(3), pp. 153–172.[22] Shamoto, E., and Moriwaki, T., 1999, “Ultaprecision Diamond Cutting of Hard-

    ened Steel by Applying Elliptical Vibration Cutting,” CIRP Ann. , 48 (1), pp.441–444.

    [23] Moriwaki, T., and Shamoto, E., 1991, “Ultraprecision Diamond Turning of Stainless Steel by Applying Ultrasonic Vibration,” CIRP Ann. , 40(1), pp.559–562.

    [24] Yan, J., Zhang, Z., and Kuriyagawa, T., 2011, “Effect of Nanoparticle Lubrica-tion in Diamond Turning of Reaction-Bonded SiC,” Int. J. Autom. Technol.,5(3), pp. 307–312. Available at: http://www.fujipress.jp/nder/xslt.php?mode=present &inputle=IJATE000500030007.xml

    [25] Inada, A., Min, S., and Ohmori, H., 2011, “Micro Cutting of Ferrous MaterialsUsing Diamond Tool Under Ionized Coolant With Carbon Particles,” CIRPAnn. , 60(1), pp. 97–100.

    [26] Yan, J., Zhang, Z., and Kuriyagawa, T., 2010, “Tool Wear Control in DiamondTurning of High-Strength Mold Materials By Means of Tool Swinging,” CIRPAnn. , 59(1), pp. 109–112.

    [27] Yoshino, M., Ogawa, Y., and Aravindan, S., 2005, “Machining of Hard-BrittleMaterials by a Single Point Tool Under External Hydrostatic Pressure,” ASMEJ. Manuf. Sci. Eng. , 127(4), pp. 837–845.

    [28] Zareena, A. R., and Veldhuis, S. C., 2012, “Tool Wear Mechanisms and ToolLife Enhancement in Ultra-Precision Machining of Titanium,” J. Mater. Pro-cess. Technol. , 212(3), pp. 560–570.

    [29] Brinksmeier, E., and Preuss, W., 2012, “Micro-Machining,” Philos. Trans. R.Soc. A , 370(1973), pp. 3973–3992.

    [30] Yan, J., Syoji, K., Kuriyagawa, T., and Suzuki, H., 2002, “Ductile RegimeTurning at Large Tool Feed,” J. Mater. Process. Technol. , 121(2–3), pp.363–372.

    [31] Kawasegi, N., Sugimori, H., Morimoto, H., Morita, N., and Hori, I., 2009, De-velopment of Cutting Tools With Microscale and Nanoscale Textures to

    Improve Frictional Behavior,” Precis. Eng. , 33(3), pp. 248–254.[32] Chang, W., Sun, J., Luo, X., Ritchie, J. M., and Mack, C., 2011, “I nvestigationof Microstructured Milling Tool for Deferring Tool Wear,” Wear , 271 (9–10),pp. 2433–2437.

    [33] Tsurimoto, S., and Moriwaki, T., 2012, “Machining of Carbides for Dies andMoulds—P 6.24,” 12th International EUSPEN conference, Stockholm.

    [34] Goel, S., Luo, X., Reuben, R. L., and Rashid, W. B., 2012, “Replacing Dia-mond Cutting Tools With CBN for Efcient Nanometric Cutting of Silicon,”Mater. Lett. , 68(0), pp. 507–509.

    [35] Tanaka, H., and Shimada, S., 2013, “Damage-Free Machining of Monocrys-talline Silicon Carbide,” CIRP Annals - Manuf. Technol. , 62(1), 2013, pp.55–58.

    [36] Yan, J., Syoji, K., and Tamaki, J., 2003, “Some Observations on the Wear of Diamond Tools in Ultra-Precision Cutting of Single-Crystal Silicon,” Wear ,255(7-12), pp. 1380–1387.

    [37] Davies, M. A., Evans, C. J., Patterson, S. R., Vohra, R., and Bergner, B. C.,2003, “Application of Precision Diamond Machining to the Manufacture of Micro-Photonics Components,” Lithographic and Micromachining Techniques for Optical Component Fabrication I I , SPIE, San Diego, CA.

    [38] Tomohiro, F., Yasuyuki, K., Takayoshi, W., Junichi, S., Kunihiro, T., and Nakai,T., 2011, “High Precision Cutting of Hardened Steel With Newly DevelopedPCBN Tools,” accessed from http://imtp.free.fr/imtp2/C4/Fukaya_Tomohiro.pdf on 10/7/2011.

    [39] Grimm, U., Muller, C., and, Wole, W. M., 2004, “Fabrication of Surfaces inOptical Quality on Pretentious Tools Steels by Ultra Precision Machining,” Pro-ceedings of the 4th Euspen International Conference, Glasgow, UK.

    [40] Patten, J., Gao, W., and Yasuto, K., 2005, “Ductile Regime Nanomachining of Single-Crystal Silicon Carbide,” ASME J. Manuf. Sci. , 127(3), pp. 522–532.

    [41] Yan, J., Zhang, Z., and Kuriyagawa, T., 2009, “Mechanism for Material Re-moval in Diamond Turning of Reaction-Bonded Silicon Carbide,” Int. J. Mach.Tools Manuf. , 49(5), pp. 366–374.

    [42] Patten, J. A., 2000, “Ductile Regime Nanocutting of Silicon Nitride,” Proceed-ings ASPE 2000 15th Annual Meeting, pp. 106–109.

    [43] Hung, N. P., and Zhong, C. H., 1996, “Cumulative Tool Wear in MachiningMetal Matrix Composites Part I: Modelling,” J. Mater. Process Technol. , 58(1),pp. 109–113.

    [44] John, A. P., Jerry, J., Biswarup, B., Andrew, G., Ning, F., and Marsh, E. R.,2007, “Chapter 2: Numerical Simulations and Cutting Experiments on Single

    Point Diamond Machining of Semiconductors and Ceramics,” Semiconductor Machining at the Micro-Nano Scale , J. Yan and J. A. Patten, eds., TransworldResearch Network, Trivandrum, Kerala, India.

    [45] Ravindra, D., and Patten, J. A., 2011, “Chapter 4: Ductile Regime Material Re-moval of Silicon Carbide(SiC),” Silicon Carbide: New Materials, Productionmethods and application , S. H. Vanger, ed., Nova Publishers, Trivandrum,India, pp. 141–167.

    [46] Niihara, K., 1979, “Slip Systems and Plastic Deformation of Silicon CarbideSingle Crystals at High Temperatures,” J. Less Common Met. , 65(1), pp.155–166.

    [47] Plimpton, S., 1995, “Fast Parallel Algorithms for Short-Range Molecular Dynamics,” J. Comput. Phys. , 117 , pp. 1–19.

    [48] Humphrey, W., Dalke, A., and Schulten, K., 1996, “VMD—Visual Molecular Dynamics,” J. Mol. Graphics , 14, pp. 33–38.

    [49] Stukowski, A., 2010, “Visualization and Analysis of Atomistic Simulation DataWith OVITO–The Open Visualization Tool,” Model. Simul. Mater. Sci. Eng. ,18(1), p. 015012.

    [50] Goel, S., 2013, “An Atomistic Investigation on the Nanometric Cutting Mecha-nism of Hard, Brittle Materials,” Mechanical Engineering , Ph.D. thesis, Heriot-

    Watt University, Edinburgh, UK, pp. 1–246.[51] Goel, S., Luo, X., Reuben, R. L., and Pen, H., 2012, “Inuence of Temperatureand Crystal Orientation on Tool Wear During Single Point Diamond Turning of Silicon,” Wear , 284–285 (0), pp. 65–72.

    [52] Goel, S., Luo, X., and Reuben, R. L., 2013, “Wear Mechanism of DiamondTools Against Single Crystal Silicon in Single Point Diamond Turning Proc-ess,” Tribol. Int. , 57(0), pp. 272–281.

    [53] Goel, S., Luo, X., Reuben, R., and Rashid, W., 2011, “Atomistic Aspects of Ductile Responses of Cubic Silicon Carbide During Nanometric Cutting,”Nanoscale Res. Lett. , 6(1), p. 589.

    [54] Notman, R., and Anwar, J., 2013, “Breaching the Skin Barrier—Insights FromMolecular Simulation of Model Membranes,” Adv. Drug Deliv. Rev. , 65(2),pp. 237–250.

    [55] Schuh, C. A., and Lund, A. C., 2003, “Atomistic Basis for the Plastic Yield Cri-terion of Metallic Glass,” Nature Mater. , 2(7), pp. 449–452.

    [56] Erhart, P., and Albe, K., 2005, “Analytical Potential for Atomistic Simulations of Silicon, Carbon, and Silicon Carbide,” Phys. Rev. B , 71(3), p. 035211.

    Journal of Manufacturing Science and Engineering APRIL 2014, Vol. 136 / 021015-11

    wnloaded From: http://manufacturingscience.asmedigitalcollection.asme.org/ on 04/14/2014 Terms of Use: http://asme.org/terms

    http://dx.doi.org/10.1177/0954405412464217http://dx.doi.org/10.1016/j.wear.2013.01.048http://dx.doi.org/10.1088/0965-0393/21/6/065004http://dx.doi.org/10.1088/0965-0393/21/6/065004http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.1063/1.4726036http://dx.doi.org/10.1016/j.ijmachtools.2012.09.001http://dx.doi.org/10.1038/479047ahttp://dx.doi.org/10.1016/j.cirp.2010.05.003http://dx.doi.org/10.1021/jp306391phttp://dx.doi.org/10.1109/IEMBS.2007.4353678http://dx.doi.org/10.1109/IEMBS.2007.4353678http://dx.doi.org/10.1109/IEMBS.2007.4353678http://dx.doi.org/10.1016/j.jmbbm.2013.12.005http://dx.doi.org/10.1016/j.jmbbm.2013.12.005http://dx.doi.org/10.1016/j.jeurceramsoc.2012.04.016http://dx.doi.org/10.1016/j.jeurceramsoc.2012.04.016http://dx.doi.org/10.1016/j.commatsci.2011.07.052http://dx.doi.org/10.1115/MSEC2009-84207http://dx.doi.org/10.1016/S0007-8506(07)62146-5http://dx.doi.org/10.1016/S0007-8506(07)62056-3http://dx.doi.org/10.1016/0141-6359(83)90063-6http://dx.doi.org/10.1016/0141-6359(83)90063-6http://dx.doi.org/10.1016/j.precisioneng.2007.08.003http://dx.doi.org/10.1016/S0007-8506(07)63222-3http://dx.doi.org/10.1016/S0007-8506(07)62053-8http://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://dx.doi.org/10.1016/j.cirp.2011.03.089http://dx.doi.org/10.1016/j.cirp.2011.03.089http://dx.doi.org/10.1016/j.cirp.2010.03.067http://dx.doi.org/10.1016/j.cirp.2010.03.067http://dx.doi.org/10.1115/1.2035695http://dx.doi.org/10.1115/1.2035695http://dx.doi.org/10.1016/j.jmatprotec.2011.10.014http://dx.doi.org/10.1016/j.jmatprotec.2011.10.014http://dx.doi.org/10.1098/rsta.2011.0056http://dx.doi.org/10.1098/rsta.2011.0056http://dx.doi.org/10.1016/S0924-0136(01)01218-3http://dx.doi.org/10.1016/j.precisioneng.2008.07.005http://dx.doi.org/10.1016/j.wear.2010.12.026http://dx.doi.org/10.1016/j.matlet.2011.11.028http://dx.doi.org/10.1016/j.matlet.2011.11.028http://dx.doi.org/10.1016/j.cirp.2013.03.098http://dx.doi.org/10.1016/S0043-1648(03)00076-0http://dx.doi.org/10.1117/12.506373http://dx.doi.org/10.1117/12.506373http://imtp.free.fr/imtp2/C4/Fukaya_Tomohiro.pdfhttp://dx.doi.org/10.1115/1.1949614http://dx.doi.org/10.1016/j.ijmachtools.2008.12.007http://dx.doi.org/10.1016/j.ijmachtools.2008.12.007http://dx.doi.org/10.1016/0924-0136(95)02114-0http://dx.doi.org/10.1016/0022-5088(79)90161-9http://dx.doi.org/10.1006/jcph.1995.1039http://dx.doi.org/10.1016/0263-7855(96)00018-5http://dx.doi.org/10.1088/0965-0393/18/1/015012http://dx.doi.org/10.1016/j.wear.2012.02.010http://dx.doi.org/10.1016/j.triboint.2012.06.027http://dx.doi.org/10.1186/1556-276X-6-589http://dx.doi.org/10.1016/j.addr.2012.02.011http://dx.doi.org/10.1038/nmat918http://dx.doi.org/10.1103/PhysRevB.71.035211http://dx.doi.org/10.1103/PhysRevB.71.035211http://dx.doi.org/10.1038/nmat918http://dx.doi.org/10.1016/j.addr.2012.02.011http://dx.doi.org/10.1186/1556-276X-6-589http://dx.doi.org/10.1016/j.triboint.2012.06.027http://dx.doi.org/10.1016/j.wear.2012.02.010http://dx.doi.org/10.1088/0965-0393/18/1/015012http://dx.doi.org/10.1016/0263-7855(96)00018-5http://dx.doi.org/10.1006/jcph.1995.1039http://dx.doi.org/10.1016/0022-5088(79)90161-9http://dx.doi.org/10.1016/0924-0136(95)02114-0http://dx.doi.org/10.1016/j.ijmachtools.2008.12.007http://dx.doi.org/10.1016/j.ijmachtools.2008.12.007http://dx.doi.org/10.1115/1.1949614http://imtp.free.fr/imtp2/C4/Fukaya_Tomohiro.pdfhttp://dx.doi.org/10.1117/12.506373http://dx.doi.org/10.1117/12.506373http://dx.doi.org/10.1016/S0043-1648(03)00076-0http://dx.doi.org/10.1016/j.cirp.2013.03.098http://dx.doi.org/10.1016/j.matlet.2011.11.028http://dx.doi.org/10.1016/j.wear.2010.12.026http://dx.doi.org/10.1016/j.precisioneng.2008.07.005http://dx.doi.org/10.1016/S0924-0136(01)01218-3http://dx.doi.org/10.1098/rsta.2011.0056http://dx.doi.org/10.1098/rsta.2011.0056http://dx.doi.org/10.1016/j.jmatprotec.2011.10.014http://dx.doi.org/10.1016/j.jmatprotec.2011.10.014http://dx.doi.org/10.1115/1.2035695http://dx.doi.org/10.1115/1.2035695http://dx.doi.org/10.1016/j.cirp.2010.03.067http://dx.doi.org/10.1016/j.cirp.2010.03.067http://dx.doi.org/10.1016/j.cirp.2011.03.089http://dx.doi.org/10.1016/j.cirp.2011.03.089http://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://www.fujipress.jp/finder/xslt.php?mode=present&hx0026;inputfile=IJATE000500030007.xmlhttp://dx.doi.org/10.1016/S0007-8506(07)62053-8http://dx.doi.org/10.1016/S0007-8506(07)63222-3http://dx.doi.org/10.1016/j.precisioneng.2007.08.003http://dx.doi.org/10.1016/0141-6359(83)90063-6http://dx.doi.org/10.1016/S0007-8506(07)62056-3http://dx.doi.org/10.1016/S0007-8506(07)62146-5http://dx.doi.org/10.1115/MSEC2009-84207http://dx.doi.org/10.1016/j.commatsci.2011.07.052http://dx.doi.org/10.1016/j.jeurceramsoc.2012.04.016http://dx.doi.org/10.1016/j.jmbbm.2013.12.005http://dx.doi.org/10.1016/j.jmbbm.2013.12.005http://dx.doi.org/10.1109/IEMBS.2007.4353678http://dx.doi.org/10.1109/IEMBS.2007.4353678http://dx.doi.org/10.1021/jp306391phttp://dx.doi.org/10.1016/j.cirp.2010.05.003http://dx.doi.org/10.1038/479047ahttp://dx.doi.org/10.1016/j.ijmachtools.2012.09.001http://dx.doi.org/10.1063/1.4726036http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.4028/www.scientific.net/KEM.496.150http://dx.doi.org/10.1088/0965-0393/21/6/065004http://dx.doi.org/10.1088/0965-0393/21/6/065004http://dx.doi.org/10.1016/j.wear.2013.01.048http://dx.doi.org/10.1177/0954405412464217

  • 8/9/2019 Surface defect machining

    12/12