Sie sind auf Seite 1von 12

Saurav Goel1

School of Mechanical
and Aerospace Engineering,
Queens University, Belfast BT95AH, UK
e-mail:.s.goel@qub.ac.uk

Waleed Bin Rashid


Institute of Mechanical, Process
and Energy Engineering,
Heriot-Watt University,
Edinburgh EH144AS, UK

Xichun Luo
Department of Design, Manufacture
and Engineering Management,
University of Strathclyde, Glasgow G11XQ, UK

Anupam Agrawal
Department of Business Administration,
University of Illinois
at Urbana Champaign, IL 61820

V. K. Jain
Department of Mechanical Engineering,
Indian Institute of Technology,
Kanpur 208016, India

A Theoretical Assessment
of Surface Defect Machining
and 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 temperature
machining at 1200 K under the same machining parameters, emulating ductile mode
micro laser assisted machining (l-LAM) and with conventional cutting at 300 K. In the
SDM 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 h100i
direction 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 plane
angle and friction coefficient 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 coefficient of friction in contrast to the isothermal cutting (carried at 1200 K)
and normal cutting (carried at 300 K). The increase of friction coefficient, 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. Analysis shows that the introduction of surface defects prior to conventional machining can be
a viable choice for machining a wide range of ceramics, hard steels and composites compared to hot machining. [DOI: 10.1115/1.4026297]
Keywords: surface defect machining, MD simulation, nanometric cutting, beta silicon
carbide

Introduction

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


method of machining that aims to obtain better quality of a
machined product at lower costs. This method utilizes predefined
and machined surface defects on the workpiece to ease material
removal. The central idea of this method is to generate surface
defects in the form of a series of holes on the top surface of the
workpiece prior to the actual machining operation. The presence
of these defects reduces the strength of the workpiece which in
turn aids to lower the cutting resistance during machining. Recent
experimental trials [1] and numerical simulations [2] on hard
steels have shown some very interesting and salient features of the
SDM method such as lower machining forces, reduction in overall
temperature in the cutting zone, reduced machining stresses and
increased chip flow velocity. SDM machining provides a product
with better surface integrity compared to that obtained using conventional hard turning.
In this work, the applicability of SDM for machining silicon
carbide is explored. Silicon carbide (SiC) is an extremely hard
and brittle nonoxide ceramic material, and its properties and manufacturing methods are being rigorously researched [36]. It has
been demonstrated that due to its superior properties, such as
chemical inertness, high thermal conductivity, high carrier saturation velocity, high specific stiffness (E/q), and high-temperature
resistance, SiC is an appropriate choice to replace silicon for
advanced ultra precision engineering applications especially in the
1
Corresponding author.
Contributed by the Manufacturing Engineering of ASME for publication in the
JOURNAL OF MANUFACTURING SCIENCE AND ENGINEERING. Manuscript received August
7, 2012; final manuscript received December 18, 2013; published online February
10, 2014. Assoc. Editor: Suhas Joshi.

electronic industry [7]. SiC is also recognized as a potential candidate for quantum computing applications as a substitute for diamond [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 in
weapons, aerospace, microelectronic and biomedical applications
as well as in big-science programmes such as the European
Extremely Large Telescope (E-ELT), the Atacama Large Millimeter/submillimeter Array (ALMA), and next generation extreme
ultraviolet (EUV) lithography steppers. SiC is also finding amazing applications in biomedical sector especially as being a semiconductor material because of being more biocompatible than
silicon [11]. Traditional orthopaedic materials such as cobalt
chrome (CoCr), stainless steel and titanium, on account of being
low wear and oxidation resistant, succumb to bone loss which
causes implant loosening resulting in a reactive implant surface.
Contrarily, SiC is capable of permanently integrating into the new
bone growth on account of low wear debris and metallosis and is
thus very effective as a coating for stents to enhance hemocompatibility and as a coating for prosthetic-bearing surfaces and
uncemented joint prosthetics [12].
In a natural state, SiC exhibits one-dimensional polymorphism:
all polytypes have the same tetrahedral arrangement of Si and C
atoms but different stacking sequences. It is due to this reason that
almost 250 polytypes of silicon carbide (SiC) have been recognized to date [13]. Across all other polytypes, two major polymorphs are a-SiC and b-SiC with hexagonal and zinc-blende
lattice structures, respectively. The main engineering properties of
b-SiC (3C-SiC) and a-SiC (6H-SiC and 4H-SiC) have already
been summarized elsewhere [14]. 3C-SiC possess extremely high
hardness and low fracture toughness and is therefore very difficult
to manufacture [15]. Considering the high hardness of 3C-SiC, the

Journal of Manufacturing Science and Engineering


C 2014 by ASME
Copyright V

APRIL 2014, Vol. 136 / 021015-1

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

Table 1 Modified form of measures suggested for improved machinability [18]


S.No.

Theoretical approach

Experimental realization

Modification of the process


1
Reduction of chemical reaction rate between diamond
cutting tool and workpiece
2
Inhibition of chemical reactions
3
Reduction of contact time between tool and workpiece
4
Lowering of temperature rise and chemical contact
5
Rotary cutting
6
Inducing external hydrostatic pressure

Cryogenic turning [19]


Use of inert gas atmospheres [20]
Vibration assisted cutting [2123]
Usage of appropriate coolant [24,25]
Tool swinging method [26]
Through externally applied hydrostatic pressure [27]

Modification of the cutting tool


6
Building a diffusion barrier on cutting tool
7
Modification of diamond lattice
8
By modifying the cutting tool geometry
9

Use of protective coatings [28]


Ion implantation [29]
Use of straight nose cutting tools [30]Providing
nanogrooves on the cutting tool [31,32].

Use of alternative cutting tool material

Using CBN instead of diamond [33,34]

Modification of the workpiece


10
Surface layer modification of the workpiece
11
Thermal softening of the workpiece during the cutting process

Ion implantation [35]


Micro laser assisted machining [17]

ing. While this approach has shown promise [17,44,45], its


adverse undesirable effects such as the lack of control on laser
power, subsurface deformation and transfer of heat to the cutting
tool has impeded its commercial realization.
Overall, the key question explored in this work is this: Can SiC
be machined using SDM method so that the machined surface
exhibits improved quality? The results from SDM machining are
compared with (i) hot machining at 1200 Ka temperature where
SiC has been reported to gain significant plasticity [46] and (ii)
normal machining at 300 K, in order to test the robustness of the
proposed method.
Fig. 1 Schematic of MD simulation model

2
hardest naturally occurring material, diamond, appears to be the
only viable choice to cut 3C-SiC. A hardness ratio of 5:1 between
the cutting tool and the workpiece is normally recommended for
machining [16], however, this ratio in the case of diamond
tool:3C-SiC workpiece is about 4:1 owing to high micro hardness
of 3C-SiC (about 2530 GPa) against diamond (100 GPa). It has
been experimentally found that nano hardness of SiC, depending
on the polytype, can vary in the range of 5570 GPa for lower
depth of cuts below 50 nm [16]. Therefore, below a cut depth less
than 50 nm, the hardness ratio of tool:workpiece reduces further to
around 2:1 and consequently the diamond tool faces tremendous
cutting resistance [16]. In such cases, methods are needed to be
identified to reduce the cutting resistance offered by the workpiece in order to improve the machinability. Microlaser assisted
machining (l-LAM) process is one such process [17], where the
workpiece is preferentially heated and thermally softened at the
tool-workpiece interface in order to improve the machinability.
Similarly, various other methods for improved machining
response of the workpiece to make them amenable to diamond
turning have been suggested over the period of time, a review is
provided in Table 1. These improvements have helped in improving the longevity of cutting tool and achieving better surface finish
of the product.
The current investigation was done on beta silicon carbide (3CSiC), but the SDM method can also be applied to other materials
which are not readily machinable through conventional
approaches, e.g., titanium [28], silicon [36], nickel [37], pretentious low carbon ferrous alloys [38,39], various polytypes of silicon carbide [40,41], silicon nitride [42] and Al-SiCp metal matrix
composites [43] where rapid tool wear and consequent deterioration to the quality of surface remains a major concern till date.
Researchers have also investigated preferential heating and thermal softening of the workpiece by a laser device prior to machin021015-2 / Vol. 136, APRIL 2014

MD Simulation Model

A public-domain computer code, known as large-scale atomic/


molecular massively parallel simulator (LAMMPS) [47] was used
to perform the MD simulation while Visual Molecular Dynamics
(VMD) [48] and OVITO [49] were used for the visualization and
analysis of the atomistic data. The following paragraphs give
details of the implementation of this code for the simulation used.
2.1 MD Simulation Model. A schematic diagram of the
nanometric cutting simulation model is shown in Fig. 1. Both the
single-crystal 3C-SiC workpiece and the diamond cutting tool
were modeled as deformable bodies in order to permit the wear
processes to take place in a natural way. The MD model developed here also incorporates a negative tool rake angle as this is
generally recommended for machining brittle materials [50].
The atoms of the substrate were allocated into one of the three
different zones: Newton atoms, thermostatic atoms and boundary
atoms, while the impacting particles were allocated to the zone of
Newton atoms. The boundary atoms were assumed to be fixed in
their initial lattice positions, serving to reduce the boundary effects
and maintain the symmetry of the lattice. Newton atoms were
allowed to follow Newtonian dynamics (LAMMPS NVE dynamics), while atoms in an intermediate thin boundary layer were subjected to a thermostat (LAMMPS NVT dynamics) to dissipate the
extra heat generated in the finite simulation volume. This consideration of boundary conditions ensures that the process of deformation is not affected by any artificial dynamics as has been done in
the past in several simulation based studies [5153]. In the case of
high temperature cutting, the temperature of both thermostatic
atoms and Newton atoms of the workpiece was scaled to 1200 K to
prevent the heat transfer between thermostatic atoms and Newton
atoms of the workpiece. It may be noted here that this work considers vacuum, which although is an unpractical consideration, but still
Transactions of the ASME

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

Fig. 2 Molecular dynamics simulation procedure (a) the initial


positions of the molecules2 are specified (b) force on each
atom due to the other atoms in its neighborhood is calculated.
(c) Potential energy function predicts the newer positions and
velocities of the atoms at specified time [54].

is a good assumption to discard the role of ambient air or gas in order


to make a careful assessment of the deformation process alone.
MD simulation (Fig. 2) is simply a classical solver of Newtons
second law of motion, where the atoms are assumed to follow
Newtons second law as follows
aix

Fix d 2 xi
2;
mi
dt

Fix 

dV
dxi

(1)

where aix represents the ith atoms acceleration in the x direction


and mi is the mass of the ith atom. Fix is the interaction force acting on the ith atom by the jth atom in the x direction, xi is ith
atoms x-coordinate and V is the potential energy function. The
force Fix acting on each particle due to the surrounding atoms is
computed from the interaction potential energy function, which is
used to determine the positions of the particles in a finite period of
time (0.5 fs in this paper). This time-stepping process is repeated
to yield a molecular trajectory which is subsequently used to analyse the process.
An area at the tool-workpiece interface which is subjected to
maximum stresses during a cutting operation is shown schematically in Fig. 3, depicting stress representation in both 3D and 2D.
One fundamental problem in the computation of atomic stress is
the fact that the volume of an atom does not remain fixed during
deformation. This makes the stress computation a challenging
task. To mitigate this problem, an elemental atomic volume is
considered in the cutting zone which is divided by the precalculated total volume of that element in order to estimate the average
stresses acting on that element during the simulation. In general,
the stress tensor for an atom i can be calculated using the following equation [47]

3
Np
Nb
1X
1X
7
6 mva vb
r1 F1b r2a F2b
r1 F1b r2a F2b
7
6
2 n1 a
2 n1 a
7
6
7
6
7
6
Na
Nd
X
X
7
6 1
1
Sab 6
r1a F1b r2a F2b r3a F3b
r1a F1b r2a F2b r3a F3b r4a F4b 7
7
63
4
7
6
n1
n1
7
6
7
6
Ni
Nf
7
6 1X
X
5
4

r1a F1b r2a F2b r3a F3b r4a F4b


ria Fib Kspaceri a; Fi b
4 n1
n1
2

where a and b denote x, y, z to generate the six components of the


symmetric stress tensor.
The first term in Eq. (2) is the contribution from the kinetic
energy of atom i. The second term is a pair-wise energy contribution where n loops over the Np neighbors of atom i and r1 and
r2 are the positions of the two atoms in the pair-wise interaction.
F1 and F2 are the forces on the two atoms resulting from the
pair-wise interactions. The third term is a bond contribution over
the Nb bonds of atom i. In a similar manner, Na angle, Nd dihedral, Ni improper interactions and Nf internal constraints of atom
i are accounted in the subsequent terms while the Kspace term
represents long-range Columbic interactions. It must be noted
here that the three body potential function such as Tersoff potential function and ABOP potential function do not include Kspace
terms in the stress computation. The above expression is however written only with an intention to comprehensively express
the mathematical formula to determine the stresses at an atomic
scale.
In a material continuum, hydrostatic stress is associated with
a change of volume leading to classical thermodynamic phase
transition, whereas the von Mises stress (corresponding to
maximum deviatoric strain energy) measures the deformation
due to shear that governs the change in shape usually by the activation of a defect transport mechanism [55], most commonly
dislocation movement. During the simulation run, the von
Mises stress (rvon Mises) was computed using the following equation [50]

2
Red and grey bodies represent two different species of atoms in a single
molecule.

Journal of Manufacturing Science and Engineering

(2)

rvon Mises
s
rxx  ryy 2 ryy  rzz 2 rzz  rxx 2 6r2xy r2yz r2zx
2
(3)
2.2. Details About the Potential Energy Function. The
potential energy function V is the backbone of any MD simulation. An example of the potential function comprising of bonded
and nonbonded interactions is shown in Fig. 4 through ball and
spring arrangements. An appropriate potential energy function
ensures the reliability of the simulation results. The selection of
potential energy function is therefore an important decision in an
MD simulation. Based on extensive tests and simulations performed prior to this work, it was found that the analytical bond
order potential (ABOP) [56] used in this work is the only available three-body potential energy function that does not suppress
the mechanism of cleavage in 3C-SiC and was therefore used in
this paper [3]. A comprehensive description along with the key
parameters of the ABOP potential function are provided elsewhere [56] in its respective reference and for the purpose of brevity they are not repeated here in this work. Table 2 lists the
computational parameters, details of the workpiece, uncut chip
thickness, details of the cutting tool and other relevant parameters
used in this study. The criterion for the selection of these parameters is driven by the fact that the author of this work has done prior
studies using these parameters and this work can therefore readily
be related to those previously published studies [52,53].
2.3 MD Simulation Setup. The MD simulation model was
developed by replicating the unit cell using periodic boundary
APRIL 2014, Vol. 136 / 021015-3

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

Fig. 3

Stresses in the cutting zone [5]

conditions along the z direction. A snapshot from the MD


simulation after equilibration process (with and without surface defects in the 3C-SiC workpiece) is shown in Figs. 5
and 6 where the red and grey colors3 correspond to silicon
and carbon atoms in the workpiece and yellow color represents carbon atoms within the diamond cutting tool,
respectively.
An important constraint of MD simulation studies is that they
are computationally expensive, and therefore use of high cutting
speeds is frequent in MD simulation studies, e.g., 5002500 m/s
cutting speed was used by Belak et al. [57,58] and Komanduri
et al. [59,60], 150400 m/s was used by Wang et al. [61] and
Liang et al. [62] and, 70100 m/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 insights
of the cutting process but in the current investigation, high cutting
speed could have affected the sensitivity of the results, particularly when cutting of the same configuration was to be compared
at 300 K and 1200 K. Therefore, current simulations were performed at a more realistic cutting speed of 10 m/s. This was
accomplished using parallel computing through MPI interface.
The calculation time for each simulation case depends on the
model size, cutting speed, cutting distance, and the number of
CPUs used.

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

021015-4 / Vol. 136, APRIL 2014

Results and Discussions

This section covers observations and discussion of the significance of the MD simulation results: cutting forces, chip morphology, stresses and temperature in the cutting zone.

Fig. 4 Potential energy function for molecular interactions in


the molecular mechanics approximation [54]

Transactions of the ASME

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

Fig. 13 Variation in the temperature of the cutting edge during


nanometric cutting of 3C-SiC

Fig. 15 Variation in shear stress acting on the tool during


nanometric cutting of 3C-SiC

to understand the process differences in detail. Figures 14 and 15


show the evolution of the von Mises stress and shear stress acting
on the cutting tool over the cutting length of 9.35 nm.
From Figs. 14 and 15, it appears that in all the three cases, the
cutting edge encounters a very high magnitude of stresses during
nanometric cutting of 3C-SiC. This can be attributed to the high
hardness of 3C-SiC, which offers tremendous cutting resistance in
comparison to other relatively softer brittle materials. A high magnitude of von Mises stress of upto 250 GPa having a component of
shear stress to the order of 125 GPa is certainly unfavorable for
the life of the diamond cutting toolthe ideal strength for diamond prior to graphitization has been reported to be around
95 GPa [66]. The stress state in all the cases suggests that graphitization of diamond tool during nanometric cutting of 3C-SiC
would be inevitable owing to the high hardness of 3C-SiC. However, SDM seems to have aided to reduce the stresses on the cutting edge to some extent.

Fig. 14 Variation in von Mises stress acting on the tool during


nanometric cutting of 3C-SiC

cutting forces intermittently and this result in minimal subsurface


damage.
3.3 Temperature and Stresses. A comparison of the evolution of the cutting temperature on the tool cutting edge in all the
three cases is shown in Fig. 13 over the cutting length of 9.35 nm.
A relatively high temperature of 480 K on the tool cutting edge is
evident during nanometric cutting at 1200 K. At higher temperatures, graphitization of diamond becomes inevitable [15,53]. The
transfer of heat from the bulk of the workpiece to the cutting tool
during hot machining may thus compromise the life of diamond
tools [15]. On the contrary, the temperature on the tool cutting
edge was observed to be much lower during SDM. This shows
that the SDM cutting process could potentially relieve the cutting
tool from high cutting loads and high temperature intermittently
which can be better for the longevity of the tool.
It will be interesting to analyse the variation in the stresses on
the cutting tool during SDM and conventional nanometric cutting
Journal of Manufacturing Science and Engineering

3.4 Surface Roughness. Experimental observations have


revealed that a significant improvement in surface roughness was
obtained when SDM was used [1]. However, it was not very clear
why improved surface roughness was achieved during SDM. Molecular dynamics simulation can help provide more accurate
understanding of such non-trivial processes. Figure 16 attempts to
address this aspect by comparing the cutting mechanism of the
SDM method with conventional machining method.
In Fig. 16, a few atoms have been colored red in order to provide some key insights into the differences between the cutting
mechanisms associated with the two cases (SDM and conventional cutting at 300 K). During conventional cutting, the atoms
are well separated during the course of machining and cling to the
cutting tool. Normally, cutting chips start flowing along the rake
face of the tool and the associated red atoms can be seen to remain
on the machined surface without contributing to any cutting
action. However, in the case of SDM, the red colored atoms4 are
concentrated below the cutting tool. Moreover, they continue to
travel with the cutting tool rather than sticking with the freshly
machined surface. This suggests that during machining these
atoms tend to rub and polish the machined surface and the local
stresses acting on the small elevated portion of the workpiece
push the particles into vacant defects, which causes the formation
4
Readers are requested to refer to the web based version of this article for correct
interpretation of the colour legends.

APRIL 2014, Vol. 136 / 021015-9

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

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

Fig. 8 Tangential cutting forces during nanometric cutting of


3C-SiC in three cases

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


three cases

cutting performed at 300 K. However, during nanometric cutting


at elevated temperature of 1200 K, both tangential cutting forces
and thrust forces reduce, albeit, to a lesser extent. On the other
hand, a noticeable reduction, especially in thrust forces, can be
seen during nanometric cutting in the case of SDM. It can be
noted here that the extent of reduction in cutting forces during
SDM 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 intermittent nature, the cutting forces and thus stresses on the cutting
tool would be relieved as soon as the cutting tool met a hole (surface defect). Table 3 summarizes various results obtained from
the simulation under different machining conditions, i.e., average
cutting forces, friction coefficient and resultant cutting forces.
The data in Table 3 show that the resultant cutting force reduces
from a value of 1449.64 nN to 1402.13 nN when the machining
was done at 1200 K instead of at 300 K, signifying a reduction in
the cutting resistance of single crystal 3C-SiC by 3.27% at
1200 K. However, the extent of this reduction is higher during the

surface defect machining process as the resultant forces drops to


1281.19 nN, i.e., a significant reduction of 11.62% compared to
normal nanometric cutting at 300 K. It is very interesting to note
here that while the cutting forces reduced, a similar trend is not
visible in the coefficient of friction. Compared to the nanometric
cutting results at 300 K, the coefficient of friction reduced by
8.11% when the cutting was performed at 1200 K. On the contrary, surface defect machining causes an increase in the friction
coefficient by 2.68%. This suggests that a different mechanism of
chip formation is associated with the proposed SDM method. This
phenomenon is discussed in detail in Sec. 3.2.
A comprehensive experimental work on surface defect machining on single crystal 3C-SiC is still underway, primarily due to the
fact that a sufficiently larger size specimen of single crystal 3CSiC is not available till date. However, experimental trials using
SDM method have already been done on AISI 4340 steel (hardened
upto 69 HRC) with diameter and depth of holes being 0.9 mm and
0.1 mm, respectively [1]. The sizes of the holes in the experimental
investigation were bigger than the size used in the MD simulation

021015-6 / Vol. 136, APRIL 2014

Transactions of the ASME

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

Table 3

Machining condition

Average tangential
cutting forces (Fc)

Average thrust
forces (Ft)

Average resultant
forces ((Ft2 Fc2))

Average friction
coefficient (Fc/Ft)

Normal machining (300 K)


Hot machining (1200 K)
Surface defect machining (300 K)

835 nN
762 nN
751 nN

1185 nN
1177 nN
1038 nN

1449.64 nN
1402.13 nN
1281.19 nN

0.7046
0.6474
0.7235

S.N.
1
2
3

Comparison of cutting forces and friction coefficient

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) and
surface defects machining (SDM)

Similarities

Differences

Cutting forces on tool

Reduced cutting forces provide better


surface finish and tool longevity.

Not applicable

Overall cutting load on tool

Not applicable

In VAM, periodic reduction in cutting load occurs at specified


amplitude, whereas in SDM cutting load reduces where
dislocations in the form of holes are encountered.

Volume of material removal

Not applicable

Although, tool is periodically rotated to reduce the cutting load, the


total 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 specified amplitude, whereas in SDM cutting chips remains in continuous contact 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 an


addition of independent process, conventional machine tool is
good enough for SDM process.

(MD cannot reach millimetre length scales due to restrictive computational speeds). However, the intent here is not to compare MD
simulation with experimental observations but to understand the
process and features of the cutting mechanism. During the experimental investigation, an improved average surface roughness (Ra)
value of 0.227 lm from SDM machining was obtained in comparison to 0.452 lm from conventional machining using a CBN tool
[1]. A more important outcome from the experiments was an intermittent reduction in the cutting forces as shown in Fig. 10.
As evident from Fig. 10, the cutting tool experiences intermittent relaxation in the cutting load during the cutting process when
surface defects are met by the cutting tool. This causes a steep
reduction in the cutting forces. This intermittent reduction in the
cutting load is favorable for tool longevity as it aids in the reduction of the local temperature at the cutting edge (we expand on
Journal of Manufacturing Science and Engineering

this later). This trend in the variation of the cutting load is more
reminiscent of vibration assisted machining and in this sense a
qualitative comparison between these two processes can be drawn
which is shown in Table 4.

3.2 Chip Morphology. Figure 11 shows a superimposed


image and a comparison of the chip morphology in all the three
cases investigated. On comparing nanometric cutting at 1200 K
with conventional nanometric cutting at 300 K, it can be seen that
the curliness of the chip has seemingly increased which is plausible 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
APRIL 2014, Vol. 136 / 021015-7

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

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

Table 5 Comparison of chip morphology and shear angle


under different machining conditions

S.N.
1
2
3

Machining
condition

Ratio of uncut chip


thickness to cut chip
thickness (r)

Shear plane
angle (u)

300 K
1200 K
SDM at 300 K

0.525
0.525
0.505

21.28 deg
21.28 deg
20.66 deg

increased and thus the shear plane angle has decreased. This was
verified using the following equation
tan /

r cos a
1  r sin a

(4)

where u is shear plane angle, a is tool rake angle, and r is the ratio
of uncut chip thickness and cut chip thickness.
Table 5 shows a decrease in shear plane angle from a value of
21.28 deg to 20.66 deg using SDM process compared to nanometric cutting at 300 K. A decrease in the value of shear plane angle
under the same machining parameters shows the dominance of
tangential cutting forces over thrust forces justifying the enhanced
cutting action of the tool. This corroborates to the increased force
ratio as seen earlier in Table 3 during the case of SDM suggesting
the dominance of tangential cutting force to be the reason of the
increase in friction coefficient (which improves the cutting
action).
Figure 12 shows the measurement of the cut chip thickness and
highlights the variation in the subsurface crystal deformation lattice layer depth. It is interesting to note that the subsurface crystal
deformation lattice layer depth becomes wider while cutting at
300 K, while it becomes a little deeper while cutting at 1200 K.
Moreover, the extent of the deformation of crystal layer underneath the finished surface is more pronounced in both these cases
compared to that in the SDM operation. It can be postulated that
high temperature weakens the bonding forces between the atoms
and hence, the atoms could easily be deformed without having
much influence on the neighbor atoms. Therefore, the deformation
did not become wider and remained concentrated under the
wake of the tool under the influence of high deviatoric stresses.
Contrarily, SDM process shows minimal subsurface deformation.
The waviness of the finished surface also seems to have decreased
during SDM. The defects generated for the purpose of SDM significantly weaken the material, which in turn reduces the bonding
strength of the atoms in the area of uncut chip thickness without
disturbing the subsurface. Also, a discontinuity in the material and
021015-8 / Vol. 136, APRIL 2014

Fig. 12 Subsurface crystal lattice deformation of 3C-SiC after


tool advances to 8.3 nm

the consequent lack of resistance to the deformation of the atoms


by the adjacent atoms makes the shearing process more preferential. Eventually, the material removal becomes easier. This is
compounded by the fact that the cutting tool is relieved from high
Transactions of the ASME

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

Fig. 13 Variation in the temperature of the cutting edge during


nanometric cutting of 3C-SiC

Fig. 15 Variation in shear stress acting on the tool during


nanometric cutting of 3C-SiC

to understand the process differences in detail. Figures 14 and 15


show the evolution of the von Mises stress and shear stress acting
on the cutting tool over the cutting length of 9.35 nm.
From Figs. 14 and 15, it appears that in all the three cases, the
cutting edge encounters a very high magnitude of stresses during
nanometric cutting of 3C-SiC. This can be attributed to the high
hardness of 3C-SiC, which offers tremendous cutting resistance in
comparison to other relatively softer brittle materials. A high magnitude of von Mises stress of upto 250 GPa having a component of
shear stress to the order of 125 GPa is certainly unfavorable for
the life of the diamond cutting toolthe ideal strength for diamond prior to graphitization has been reported to be around
95 GPa [66]. The stress state in all the cases suggests that graphitization of diamond tool during nanometric cutting of 3C-SiC
would be inevitable owing to the high hardness of 3C-SiC. However, SDM seems to have aided to reduce the stresses on the cutting edge to some extent.

Fig. 14 Variation in von Mises stress acting on the tool during


nanometric cutting of 3C-SiC

cutting forces intermittently and this result in minimal subsurface


damage.
3.3 Temperature and Stresses. A comparison of the evolution of the cutting temperature on the tool cutting edge in all the
three cases is shown in Fig. 13 over the cutting length of 9.35 nm.
A relatively high temperature of 480 K on the tool cutting edge is
evident during nanometric cutting at 1200 K. At higher temperatures, graphitization of diamond becomes inevitable [15,53]. The
transfer of heat from the bulk of the workpiece to the cutting tool
during hot machining may thus compromise the life of diamond
tools [15]. On the contrary, the temperature on the tool cutting
edge was observed to be much lower during SDM. This shows
that the SDM cutting process could potentially relieve the cutting
tool from high cutting loads and high temperature intermittently
which can be better for the longevity of the tool.
It will be interesting to analyse the variation in the stresses on
the cutting tool during SDM and conventional nanometric cutting
Journal of Manufacturing Science and Engineering

3.4 Surface Roughness. Experimental observations have


revealed that a significant improvement in surface roughness was
obtained when SDM was used [1]. However, it was not very clear
why improved surface roughness was achieved during SDM. Molecular dynamics simulation can help provide more accurate
understanding of such non-trivial processes. Figure 16 attempts to
address this aspect by comparing the cutting mechanism of the
SDM method with conventional machining method.
In Fig. 16, a few atoms have been colored red in order to provide some key insights into the differences between the cutting
mechanisms associated with the two cases (SDM and conventional cutting at 300 K). During conventional cutting, the atoms
are well separated during the course of machining and cling to the
cutting tool. Normally, cutting chips start flowing along the rake
face of the tool and the associated red atoms can be seen to remain
on the machined surface without contributing to any cutting
action. However, in the case of SDM, the red colored atoms4 are
concentrated below the cutting tool. Moreover, they continue to
travel with the cutting tool rather than sticking with the freshly
machined surface. This suggests that during machining these
atoms tend to rub and polish the machined surface and the local
stresses acting on the small elevated portion of the workpiece
push the particles into vacant defects, which causes the formation
4
Readers are requested to refer to the web based version of this article for correct
interpretation of the colour legends.

APRIL 2014, Vol. 136 / 021015-9

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

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

of a layer of quasi-continuous material. Particles of this material


trapped below the tool continue to rub the machined surface and
destroy the remains of the feed marks left by the cutting tool.
Therefore, it appears that in addition to the SDM machining
mechanism occurring at the rake face, a simultaneous polishing
mechanism also proceeds with the travel of the cutting tool.

Conclusions

A new machining approach named surface defect machining


(SDM) for machining single crystal 3C-SiC at nanoscale has been
tested using MD simulation and compared with ductile mode
micro laser assisted machining at 1200 K and conventional
machining at 300 K. The motivation of this newly developed
method was the anticipation that the surface defects generated on
the workpiece can allow easy shearing of the material. The comprehensive results obtained from the simulation can be concluded
as follows:
(1) The presence of premachined surface defects improves the
machinability of difficult-to-machine materials through a
reduction in shear plane angle and shear plane area thus permitting reduced side flow with less metallurgical transformations on the finished machined surface and the subsurface.
(2) Surface defects cause a reduction in the average cutting
force which relieves the cutting tool of the cutting load
021015-10 / Vol. 136, APRIL 2014

intermittently. This was attributed to the lowered bonding


strength between the workpiece atoms. While this process
is reminiscent of a vibration assisted machining process,
the added advantage is that it does not require any separate
attachment.
(3) The extent of subsurface crystal deformation lattice layer
depth was found to be minimal in the case of surface defect
machining followed by high temperature cutting whereas
cutting performed at 300 K showed maximum depth of the
subsurface deformation. A relatively higher temperature on
the tool cutting edge was found as an inherent characteristic
with the l-LAM process in contrast to SDM process. Consequently, l-LAM process presents the risk of accelerated
graphitization of the diamond tools. This is because high
temperature nanometric cutting causes transfer of heat from
the bulk of the workpiece to the cutting tool which may
compromise the life of the diamond tools. On the contrary,
SDM helps to reduce the temperature at the tool cutting
edge compared to the high temperature cutting and the normal nanometric cutting at 300 K.

Acknowledgment
Authors acknowledge the funding support of Ministry of Higher
Education, Kingdom of Saudi Arabia for funding the Ph.D of
Transactions of the ASME

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

W.B.R. and an additional funding from J. M. Lessells travel


scholarship from the Royal Society of Edinburgh (2013 RSE/J. M.
Lessells Travel Scholarship), International Research Fellowship
account of Queens University, Belfast and an EPSRC research
grant (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 Roughness
During Hard Turning Process, Proc. Inst. Mech. Eng., Part B, 227(2), pp.
338342.
[2] Rashid, W. B., Goel, S., Luo, X., and Ritchie, J. M., 2013, The Development
of a Surface Defect Machining Method for Hard Turning Processes, Wear,
302(12), pp. 11241135.
[3] Goel, S., Stukowski, A., Luo, X., Agrawal, A., and Reuben, R. L., 2013,
Anisotropy of Single-Crystal 3CSiC 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, Single
Point Diamond Turning of Single Crystal Silicon Carbide: Molecular Dynamic
Simulation Study, Key Engineering Materials, Liverpool. pp. 150155.
[5] Goel, S., Luo, X., and Reuben, R. L., 2012, Shear Instability of Nanocrystalline 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,
BrittleDuctile Transition During Diamond Turning of Single Crystal Silicon
Carbide, Int. J. Mach. Tools Manuf., 65, pp. 1521.
[7] Neudeck, P. G., 2000, SiC Technology, The VLSI Handbook, B. Raton, ed.,
CRC Press , Boca Raton, FL, pp. 6.16.24.
[8] Dzurak, A., 2011, Quantum Computing: Diamond and Silicon Converge,
Nature, 479(7371), pp. 4748.
[9] Shore, P., Cunningham, C., DeBra, D., Evans, C., Hough, J., Gilmozzi, R.,
Kunzmann, H., Morantz, P., and Tonnellier, X., 2010, Precision Engineering
for Astronomy and Gravity Science, CIRP Ann., 59(2), pp. 694716.
[10] Newsome, D. A., Sengupta, D., Foroutan, H., Russo, M. F., and van Duin, A.
C., 2012, Oxidation of Silicon Carbide by O2 and H2O: A ReaxFF Reactive
Molecular Dynamics Study, Part I, J. Phys. Chem. C, 116(30), pp.
1611116121.
[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 Medicine
and Biology Society, EMBS IEEE.
[12] Goel, S., Yan, J., Luo, X., and Agrawal, A., 2013, Incipient Plasticity in 4HSiC During Quasistatic Nanoindentation, J. Mech. Behav. of Biomed.l Materials, (in press).
[13] Perrone, D., 2007, Process and Characterisation Techniques on 4HSilicon
Carbide, 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 Silicon
Carbide, J. Europ. Ceram. Soc., 32(12), pp. 34233434.
[15] Goel, S., Luo, X., and Reuben, R. L., 2012, Molecular Dynamics Simulation
Model for the Quantitative Assessment of Tool Wear During Single Point Diamond Turning of Cubic Silicon Carbide, Comput. Mater. Sci., 51(1), pp.
402408.
[16] Ravindra, D., 2011, Ductile Mode Material Removal of Ceramics and Semiconductors, 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 (lLAM) Process ASME Conf. Proc., MSEC2009 -84207, pp. 827832.
[18] Brinksmeier, E., and Glabe, R., 2001, Advances in Precision Machining of
Steel, CIRP Ann., 50(1), pp. 385388.
[19] Evans, C., and Bryan, J. B., 1991, Cryogenic Diamond Turning of Stainless
Steel, CIRP Ann., 40(1), pp. 571575.
[20] Casstevens, J. M., 1983, Diamond Turning of Steel in Carbon-Saturated
Atmospheres, Precis. Eng., 5(1), pp. 915.
[21] Brehl, D. E., and Dow, T. A., 2008, Review of Vibration-Assisted Machining,
Precis. Eng., 32(3), pp. 153172.
[22] Shamoto, E., and Moriwaki, T., 1999, Ultaprecision Diamond Cutting of Hardened Steel by Applying Elliptical Vibration Cutting, CIRP Ann., 48(1), pp.
441444.
[23] Moriwaki, T., and Shamoto, E., 1991, Ultraprecision Diamond Turning of
Stainless Steel by Applying Ultrasonic Vibration, CIRP Ann., 40(1), pp.
559562.
[24] Yan, J., Zhang, Z., and Kuriyagawa, T., 2011, Effect of Nanoparticle Lubrication in Diamond Turning of Reaction-Bonded SiC, Int. J. Autom. Technol.,
5(3), pp. 307312. Available at: http://www.fujipress.jp/finder/xslt.php?mode=
present&inputfile=IJATE000500030007.xml
[25] Inada, A., Min, S., and Ohmori, H., 2011, Micro Cutting of Ferrous Materials
Using Diamond Tool Under Ionized Coolant With Carbon Particles, CIRP
Ann., 60(1), pp. 97100.
[26] Yan, J., Zhang, Z., and Kuriyagawa, T., 2010, Tool Wear Control in Diamond
Turning of High-Strength Mold Materials By Means of Tool Swinging, CIRP
Ann., 59(1), pp. 109112.

Journal of Manufacturing Science and Engineering

[27] Yoshino, M., Ogawa, Y., and Aravindan, S., 2005, Machining of Hard-Brittle
Materials by a Single Point Tool Under External Hydrostatic Pressure, ASME
J. Manuf. Sci. Eng., 127(4), pp. 837845.
[28] Zareena, A. R., and Veldhuis, S. C., 2012, Tool Wear Mechanisms and Tool
Life Enhancement in Ultra-Precision Machining of Titanium, J. Mater. Process. Technol., 212(3), pp. 560570.
[29] Brinksmeier, E., and Preuss, W., 2012, Micro-Machining, Philos. Trans. R.
Soc. A, 370(1973), pp. 39733992.
[30] Yan, J., Syoji, K., Kuriyagawa, T., and Suzuki, H., 2002, Ductile Regime
Turning at Large Tool Feed, J. Mater. Process. Technol., 121(23), pp.
363372.
[31] Kawasegi, N., Sugimori, H., Morimoto, H., Morita, N., and Hori, I., 2009, Development of Cutting Tools With Microscale and Nanoscale Textures to
Improve Frictional Behavior, Precis. Eng., 33(3), pp. 248254.
[32] Chang, W., Sun, J., Luo, X., Ritchie, J. M., and Mack, C., 2011, Investigation
of Microstructured Milling Tool for Deferring Tool Wear, Wear, 271(910),
pp. 24332437.
[33] Tsurimoto, S., and Moriwaki, T., 2012, Machining of Carbides for Dies and
MouldsP 6.24, 12th International EUSPEN conference, Stockholm.
[34] Goel, S., Luo, X., Reuben, R. L., and Rashid, W. B., 2012, Replacing Diamond Cutting Tools With CBN for Efficient Nanometric Cutting of Silicon,
Mater. Lett., 68(0), pp. 507509.
[35] Tanaka, H., and Shimada, S., 2013, Damage-Free Machining of Monocrystalline Silicon Carbide, CIRP Annals - Manuf. Technol., 62(1), 2013, pp.
5558.
[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. 13801387.
[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 II, 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 Developed
PCBN Tools, accessed from http://imtp.free.fr/imtp2/C4/Fukaya_Tomohiro.pdf
on 10/7/2011.
[39] Grimm, U., Muller, C., and, Wolfle, W. M., 2004, Fabrication of Surfaces in
Optical Quality on Pretentious Tools Steels by Ultra Precision Machining, Proceedings of the 4th Euspen International Conference, Glasgow, UK.
[40] Patten, J., Gao, W., and Yasuto, K., 2005, Ductile Regime Nanomachining of SingleCrystal Silicon Carbide, ASME J. Manuf. Sci., 127(3), pp. 522532.
[41] Yan, J., Zhang, Z., and Kuriyagawa, T., 2009, Mechanism for Material Removal in Diamond Turning of Reaction-Bonded Silicon Carbide, Int. J. Mach.
Tools Manuf., 49(5), pp. 366374.
[42] Patten, J. A., 2000, Ductile Regime Nanocutting of Silicon Nitride, Proceedings ASPE 2000 15th Annual Meeting, pp. 106109.
[43] Hung, N. P., and Zhong, C. H., 1996, Cumulative Tool Wear in Machining
Metal Matrix Composites Part I: Modelling, J. Mater. Process Technol., 58(1),
pp. 109113.
[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., Transworld
Research Network, Trivandrum, Kerala, India.
[45] Ravindra, D., and Patten, J. A., 2011, Chapter 4: Ductile Regime Material Removal of Silicon Carbide(SiC), Silicon Carbide: New Materials, Production
methods and application, S. H. Vanger, ed., Nova Publishers, Trivandrum,
India, pp. 141167.
[46] Niihara, K., 1979, Slip Systems and Plastic Deformation of Silicon Carbide
Single Crystals at High Temperatures, J. Less Common Met., 65(1), pp.
155166.
[47] Plimpton, S., 1995, Fast Parallel Algorithms for Short-Range Molecular
Dynamics, J. Comput. Phys., 117, pp. 119.
[48] Humphrey, W., Dalke, A., and Schulten, K., 1996, VMDVisual Molecular
Dynamics, J. Mol. Graphics, 14, pp. 3338.
[49] Stukowski, A., 2010, Visualization and Analysis of Atomistic Simulation Data
With OVITOThe Open Visualization Tool, Model. Simul. Mater. Sci. Eng.,
18(1), p. 015012.
[50] Goel, S., 2013, An Atomistic Investigation on the Nanometric Cutting Mechanism of Hard, Brittle Materials, Mechanical Engineering, Ph.D. thesis, HeriotWatt University, Edinburgh, UK, pp. 1246.
[51] Goel, S., Luo, X., Reuben, R. L., and Pen, H., 2012, Influence of Temperature
and Crystal Orientation on Tool Wear During Single Point Diamond Turning of
Silicon, Wear, 284285(0), pp. 6572.
[52] Goel, S., Luo, X., and Reuben, R. L., 2013, Wear Mechanism of Diamond
Tools Against Single Crystal Silicon in Single Point Diamond Turning Process, Tribol. Int., 57(0), pp. 272281.
[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 BarrierInsights From
Molecular Simulation of Model Membranes, Adv. Drug Deliv. Rev., 65(2),
pp. 237250.
[55] Schuh, C. A., and Lund, A. C., 2003, Atomistic Basis for the Plastic Yield Criterion of Metallic Glass, Nature Mater., 2(7), pp. 449452.
[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.

APRIL 2014, Vol. 136 / 021015-11

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

[57] Belak, J. F., and Stowers, I. F., 1990, A Molecular Dynamics model of Orthogonal Cutting Process, Proceedings of American Society Precision Engineering
Annual conference, pp. 7679.
[58] Belak, J., 1994, Nanotribology: Modelling Atoms When Surfaces Collide, Energy
and Technology Review, Lawrence Livermore National Laboratory, Livermore, CA.
[59] Komanduri, R., Chandrasekaran, N., and Raff, L. M., 2000, Molecular Dynamics
Simulation of Atomic-Scale Friction, Phys. Rev. B, 61(20), pp. 1400714019.
[60] Komanduri, R., Chandrasekaran, N., and Raff, L. M., 2000, Simulation of
Nanometric Cutting of Single Crystal AluminumEffect of Crystal Orientation
and Direction of Cutting, Wear, 242(12), pp. 6088.
[61] Wang, Z., Liang, Y., Chen, M., Tong, Z., and Chen, J., 2010, Analysis About
Diamond Tool Wear in Nano-Metric Cutting of Single Crystal Silicon Using
Molecular Dynamics Method, SPIE Bellingham, WA.

021015-12 / Vol. 136, APRIL 2014

[62] Liang, Y. C., Guo, Y. B., Chen, M. J., and Bai, Q. S., 2008, Molecular Dynamics Simulation of Heat Distribution During Nanometric Cutting Process, 2nd
IEEE International of Nanoelectronics Conference, INEC 2008.
[63] Noreyan, A., Amar, J. G., and Marinescu, I., 2005, Molecular Dynamics Simulations of Nanoindentation of Beta-SiC With Diamond Indenter, Mater. Sci.
Eng. B, 117(3), pp. 235240.
[64] Noreyan, A., and Amar, J. G., 265, 2008, Molecular Dynamics Simulations of
Nanoscratching of 3C SiC, Wear, 265(7-8), pp. 956962.
[65] Rentsch, R., and Inasaki, I., 2006, Effects of Fluids on the Surface Generation
in Material Removal ProcessesMolecular Dynamics Simulation, CIRP
Ann., 55(1), pp. 601604.
[66] Roundy, D., and Cohen, M. L., 2001, Ideal Strength of Diamond, Si, and Ge,
Phys. Rev. B, 64(21), p. 212103.

Transactions of the ASME

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

Das könnte Ihnen auch gefallen