You are on page 1of 29

Wear 242 Ž2000.

60–88
www.elsevier.comrlocaterwear

M.D. Simulation of nanometric cutting of single crystal


aluminum–effect of crystal orientation and direction of cutting
R. Komanduri ) , N. Chandrasekaran, L.M. Raff 1
Mechanical & Aerospace Engineering, Oklahoma State UniÕersity, 218 Engineering North, Stillwater, OK 74078 USA
Received 7 October 1998; received in revised form 8 March 2000; accepted 24 March 2000

Abstract

Molecular Dynamics ŽMD. simulation of nanometric cutting was conducted on single crystal aluminum in specific combinations of
crystal orientation Ž111., Ž110., and Ž001.4 and cutting direction ²w110x, w211x, and w100x: and with tools of different positive rake angles
Ž08, 108, and 408. to investigate the nature of deformation and the extent of anisotropy of this material. When the aluminum crystal was
oriented in Ž111. plane and cut in w110x direction, plastic deformation ahead of the tool was accomplished predominantly by compression
along with shear in the cutting direction. Also, the deformation in the work material Žunderneath the depth of cut region. was found to be
along the cutting direction. In Ž001. w110x combination, the dislocations were found to be generated parallel to the cutting direction. These
were relieved from the uncut region into the work material underneath by elastic recovery. While there was some reorganization, yet
some disorder of the atoms was observed in the machined surface in the amount close to the depth of cut. In contrast, in Ž110. w001x
combination, the dislocations were generated normal to the cutting direction, which was rather unusual in machining. In the case of Ž110.
orientation and w110x cutting direction, the dislocations were found to be parallel as well as perpendicular to the cutting direction. In
contrast, for Ž001. w100x combination, extensive dislocations motion at ; 458 to the cutting direction was seen. Similarly, for Ž111. w211x
combination, the dislocation motion was observed to be at ; 608 to the cutting direction. In both cases, the material in the shear zone was
deformed at an angle Žequivalent to a shear angle., which is the mirror image of the dislocations generated in the work material, i.e.,
; 458 Žclockwise. in the case of Ž001. orientation and w100x cutting direction and ; 608 Žclockwise. in the case of Ž111. orientation and
w211x cutting direction. The variation of the cutting forces, the ratio of thrust to cutting force, the specific energy Ženergy required for
removal of unit volume of work material., and the nature of deformation ahead of the tool as well as the subsurface deformation of the
machined surface with crystal orientation and direction of cutting were investigated. q 2000 Elsevier Science S.A. All rights reserved.

Keywords: Molecular dynamics simulation; Nanometric cutting; Single crystal aluminum

1. Introduction process is termed nanometric cutting. In the case of ultra-


precision machining of conventional polycrystalline mate-
Precision aluminum disks are used extensively for hard rials Ži.e., with a grain size of several micrometers., as the
drives in the computer industry and precision aluminum depth of cut is an extremely small fraction of the average
mirrors are used in numerous optical applications as in grain size, UPM basically involves cutting within a single
lasers and rotating mirrors in copying machines. Produc- grain or single crystal with periodic interruptions at the
tion of these devices involve ultraprecision machining grain boundaries. Since grain orientation changes from one
ŽUPM. of a single crystal or polycrystalline work material crystal to another in a polycrystalline aggregate, albeit
at fine depths of cut Žon the order of a few nanometers or small, the cutting tool in nanometric cutting experiences
less. with a single crystal diamond tool. Consequently, this work material with different crystallographic orientations
and directions of cutting. However, there is hardly any
information on the effect of crystal orientation and direc-
) tion of cutting aluminum work material. In addition, there
Corresponding author. Tel.: q1-405-744-5900; fax: q1-405-744-
7873.
are many precision engineering applications where single
E-mail address: ranga@ceat.okstate.edu ŽR. Komanduri.. crystal materials like silicon, germanium are finished by
1
Chemistry Department. nanometric cutting.

0043-1648r00r$ - see front matter q 2000 Elsevier Science S.A. All rights reserved.
PII: S 0 0 4 3 - 1 6 4 8 Ž 0 0 . 0 0 3 8 9 - 6
R. Komanduri et al.r Wear 242 (2000) 60–88 61

Since, some of the single crystal materials are Table 1


anisotropic in their physical and mechanical properties, it Summary of elastic properties of polycrystalline aluminum as well as the
stiffness, compliance constants and elastic anisotropy of single crystal
is appropriate to investigate their behavior in nanometric aluminum in different directions
cutting, namely, the variation of the cutting forces, me-
Elastic modulus ŽGPa., Polycrystalline 70.30
chanics of chip formation, specific energy, etc., in different Shear Modulus ŽGPa., Polycrystalline 26.1
crystal orientations and cutting directions. Consequently, Poisson’s ratio, Polycrystalline 0.345
the nanometric cutting of single crystal aluminum is of Burger’s Vector 2.86
great technological significance to the computer, optical, Relative degree of anisotropy a 1.219
laser, and printing industries in addition to its scientific E 111 ŽGPa. 76.1
E 100 ŽGPa. 63.7
relevance. In this paper, the results of the Molecular E 110 ŽGPa. 72.59
Dynamics ŽMD. simulation of nanometric cutting of single E 111 rE 100 1.19
crystal aluminum in specific combinations of crystal orien- C 11Ž10 10 Pa. 10.82
tation and cutting direction with tools of different rake C 12 Ž10 10 Pa. 6.13
angles are reported to investigate the anisotropic behavior C 44 Ž10 10 Pa. 2.85
S11Ž10y11 Pay1 . 1.57
of this material. S12 Ž10y11 Pay1 . y0.57
Material removal at small depths of cut involves plastic S 44 Ž10y11 Pay1 . 3.51
deformation ahead of the tool and elastic recovery of the
a
finished surface Žmachined surface underneath the clear- 2ŽS11 yS12 .rS 44 .
ance face of the tool.. The elastic recovery of the ma-
chined surface results in small upward bulge of material at
the machined surface near the tool tip and consequent
rubbing of it with the clearance face of the tool. In For example, it is well known that for the FCC metals, slip
conventional cutting where the depth of cut is significant, is predominant on the  1114 planes and w110x directions. In
the elastic effects are rather small compared to the plastic nanometric cutting one would, therefore, expect these
deformation effects and hence may be neglected. This is planes and directions to result in easy slip and perhaps
not the case in nanometric cutting due to extremely small lower cutting forces. It would be interesting to investigate
depths of cut, on the order of a few nanometers or less. the degree of anisotropy of aluminum in nanometric cut-
Thus, the variation in the elastic modulus with crystallo- ting and compare it with its value in the elastic region,
graphic orientation can significantly affect the finish, accu- which is 1.219. Variations in the plastic behavior of the
racy, and surface integrity of the machined components. material in cutting with different crystallographic orienta-
Consequently, the accuracy of the machining process and tions can thus be determined.
the quality of the finished surface Žsurface integrity. de-
pend upon both the plastic and elastic behavior of the
material being machined. 2. Literature review
Since the binding forces are strongly affected by the
distance between the atoms in a crystal, the elastic con- 2.1. Machining of single crystal materials
stants of a single crystal will vary with the direction. For
example, the modulus of elasticity in any direction of a Using ultramicrotomy, Black w3x investigated the nature
cubic crystal is given by w1x: of plastic deformation of single crystal aluminum in ma-
chining at fine depths of cut. He reported the discontinuous
1 1 nature of the machining process on a microscale leading to
s S11 y 2 Ž S11 y S12 . y S44 the systematic formation of ‘‘shear lamellae’’ separated by
E 2 ‘‘shear fronts.’’ He observed large variations in the chip
thickness ratio Žratio of chip thickness to the uncut chip
= Ž l 2 m2 q m2 n2 q l 2 n2 . , thickness. with variation in the crystal orientation. Rama-
lingam and Hazra w4x conducted cutting experiments on
where Si j are the elastic constants in different orientations single crystal aluminum of known crystal orientations.
and l, m, and n are the direction cosines. They found the dynamic shear stress ŽDSS. to remain
Table 1 gives a summary of the elastic properties of constant for all orientations and consequently concluded
polycrystalline and single crystal aluminum in different the DSS to be a material property. They also reported a
directions Žstiffness, compliance constants, and elastic ani- wide variation in the shear angle with orientation. Several
sotropy.. It may be noted that aluminum has a degree of other research workers have suggested the constancy of
anisotropy of 1.219 in the elastic region. In comparison, DSS and attributed DSS as a true material property w5–7x.
copper has been reported to have much higher degree of Williams and Gane w8x conducted orthogonal cutting
anisotropy of 3.203 w2x. A somewhat similar anisotropy in experiments on single crystal copper in two orientations
the plastic range can also be anticipated of these materials. — one with the Ž111. plane parallel to the slip plane and
62 R. Komanduri et al.r Wear 242 (2000) 60–88

the other parallel to the cutting direction. In contrast to the Moriwaki et al. w14x conducted in situ machining exper-
results reported by Ramalingam and Hazra w4x, they ob- iments inside the SEM on a single crystal copper in Ž110.
served the shear angle to be relatively unchanged and the and Ž111. planes and in various cutting directions at a
resolved shear stress ŽRSS. dependent on the orientation. depth of cut ranging from 0.1 to 5 mm and a cutting speed
They also observed the shear stress to be ; 40% lower of 120 mmrmin. They found the crystallographic orienta-
when Ž111. plane was parallel to the shear plane as this tion affects the chip formation process in terms of magni-
orientation favored easy chip formation. In a subsequent tude of the shear angle and the cutting force. The shear
work, Williams w9x observed copper to display a depen- angle was found to reach as high as 608. Moriwaki et al.
dence of RSS with orientation while aluminum did not, w14x observed that the mechanics of chip formation was not
thus validating the results of both Ramalingam and Hazra much influenced as the depth of cut was reduced from
w4x and Williams and Gane w8x. He attributed this differ- conventional to micrometer Žor even submicrometer. depths
ence to the variation in the nature of dislocation motion in of cut. Moriwaki et al. w15x in a subsequent study on
aluminum as compared to copper due to differences in the ultraprecision orthogonal cutting of single crystal copper
stacking fault energy and equilibrium spacing of the atoms with a diamond tool, concluded that the cutting forces,
in the crystal. shear angle, and surface roughness are not much influ-
Ueda and Iwata w10x investigated the mechanism of chip enced by the crystal orientation when the depth of cut is
formation, variation of cutting forces, and shear angle with decreased to submicrometer level.
the crystallographic orientation during diamond cutting of
b-brass. They observed typical lamella slip structure and
reported the formation of discontinuous chip in a particular 2.2. MD simulation studies
range of crystal orientations. They observed the shear
angle to vary from ; 158 to 608 with changes in crystallo- MD simulation studies, in general, were initiated in the
graphic orientations. These measured values were found to late 1950’s by Alder and Wainwright w16,17x in the field of
be in reasonable agreement with the estimated values statistical mechanics. Since then, MD simulation has been
based on Schmid factor. The cutting forces and surface applied to a wide range of fields including crystal growth,
roughness values were also observed to depend on the indentation, tribology, low-pressure diamond synthesis, and
material anisotropy. However, no explanation was offered laser interactions, to name a few w18–27x. However, its
for these unusually high shear angles. application to machining is only of recent origin. For
In most theoretical analyses of the cutting process, the example, in the late 1980’s, a group at Lawrence Liver-
work material is assumed to be isotropic and homogeneous more National Laboratories ŽLLNL. had conducted pio-
for convenience. Lee w11x proposed a physical model to neering studies on MD simulation of nanometric cutting of
predict the variation of cutting forces due to different copper with a diamond tool w28–34x. This work led other
crystallographic orientations based on the changes in the researchers, particularly from Japan, and in the author’s
shear angle of the crystal being machined. In a subsequent laboratory from the U.S. to explore and extend MD simu-
study, Lee and Zhou w12x proposed a microplasticity model lation of nanometric cutting to many practical machining
to analyze the effect of crystallographic orientation on the applications. In this connection, the work of Ikawa et al.
shear zone formation in micromachining. They proposed w35–38x, Shimada w39x, and Inamura et al. w40–43x of
that the most likely shear angle is the one that has the most Japan are particularly noteworthy.
negative texture softening factor among the ones with the Belak et al. w31,32x studied both two-dimensional Ž2-D.
same minimum shear strength. Based on this, they con- and three-dimensional Ž3-D. cutting of copper using the
cluded that the variation in microcutting force can be embedded-atom method. Belak w33x also investigated the
predicted if the change in crystallographic orientation of machining of silicon using a diamond tool. Belak and
the substrate material with respect to the cutting direction Stowers w34x generated computer animated movies of the
is known. MD simulation of nanometric cutting, which are very
Konig and Senrath w13x conducted cutting experiments informative. Ikawa et al. w35–38x and Shimada w39x studied
on monocrystalline copper ŽOFHC. substrate with  1004 , the 2-D nanometric cutting of copper with a diamond tool.
 1104 , and  1114 -oriented crystals along different cutting They investigated the effect of edge radius and minimum
directions. They observed a significant dynamic compo- depth of cut on the chip formation process, subsurface
nent of the cutting force along ²100: cutting direction, deformation, and specific cutting energy in two cutting
which they attributed to the resulting poor surface quality directions of the workpiece, namely ²110: and ²121:.
of the machined surface. However, when the substrate was Inamura et al. w40–43x reported MD simulation under
machined along ²110: direction, the dynamic component quasi-static conditions where only the change in the mini-
of the cutting force was low and consequently, a high mum-energy positions, which are the mean positions of the
surface quality was obtained. They also predicted the vibrating atoms were followed. Inamura et al. w43x pre-
plastic material properties to be the cause for the surface sented a method of transformation from an atomic model
slip encountered at the grain boundaries. Žnanometric cutting. to an equivalent continuum ŽFEM.
R. Komanduri et al.r Wear 242 (2000) 60–88 63

model Žmachining.. Following the work of Belak et al


w31,32x, Chandrasekaran et al. w44x investigated a method
termed the Length Restricted Molecular Dynamics ŽLRMD.
simulation to reduce the computational time and at the
same time reduce the memory requirements significantly.
Komanduri et al. w45x reported MD simulation studies of
machining with large negative rake angle tools to simulate
grinding and compared the simulation results with the
experimental results resulting in good agreement. Koman-
duri et al. w46x also investigated the effect of tool geometry
with tools of different edge radii relative to the depth of
cut in nanometric cutting. Overall, while a sound founda-
tion has been laid for the MD simulation of nanometric
Fig. 1. Schematic of the model used in the MD simulation of nanometric
cutting, many of the challenging machining problems are
cutting of single crystal aluminum to investigate the effect of crystallo-
yet to be investigated. graphic orientation and direction of cutting. Note that the potential
Only limited results in the ultraprecision machining of includes no interaction terms between fixed atoms.
single crystal aluminum in different orientations have been
reported in the literature. Sato et al. w47x conducted micro-
cutting experiments on aluminum single crystals oriented symmetrically placed about the origin, then the boundary
along different crystallographic planes. They reported a atoms are located on the qX, yX, qY, yY, and yZ
better surface finish when cutting along w011x direction as planes of the work material. The qZ side corresponds to
compared to w121x direction. To et al. w48x conducted the top surface of the work material where the chip is
diamond turning of aluminum single crystal rods with being formed and hence no boundary atoms are placed
crystallographic axes normal to ²100:, ²110:, and ²111: along this face. The height Ž Z-axis. w10.13 nmx and length
using a cutting tool of 0.635 mm edge radius, 08 rake angle Ž Y-axis. w12.15 nmx of the work material were selected
and 58 clearance angle and depths of cut in the range of large enough to avoid or minimize the boundary effects for
1–10 mm. Continuous chip formation was observed under the simulated depth of cut w0.81–1.62 nmx, but small
all cutting conditions. They reported highest cutting force enough to be able to simulate in reasonable amount of
for the  1104 -oriented crystals and lowest for the  1114 -ori- time. The width of the work material Ž X-axis. w1.62 nmx is
ented crystals. They also observed the Ž001.-oriented crys- small in comparison to the height and length of the work
tals to exhibit a relatively low magnitude of the dynamic material. Consequently, it may have some effect on the
component of the force compared to Ž110.- or Ž111.-ori- mechanics of the process. But such a choice is necessary to
ented crystals. To et al. w48x suggest a high surface finish minimize the computational time, which would otherwise
with Ž001. plane but a poor surface finish when turning take several weeks for each run. It will be shown in the
was carried out on Ž110. plane. According to the authors, following that a change in the width of the work material
the depth of cut had negligible effect on the surface does not affect the process mechanics Žshear angle, dislo-
roughness over the range of depth of cut Ž1–10 mm. cation orientation, etc.. or the results studied in the current
studied in their experiments. investigation Žforces, force ratio, and specific energy. sig-
nificantly. Further, the objective here is to simulate orthog-
onal or 2-D cutting under plane strain conditions. Also by
3. Methodology for MD simulation of nanometric cut- increasing the width, the number of layers on the MD
ting simulation plots will increase and this will make the
delineation of the deformation mechanism somewhat diffi-
cult to analyze.
3.1. Cutting model The motion of the atoms in the moving zone is deter-
mined solely by the forces produced by the interaction
Fig. 1 is a schematic of the model used Žshown in potential and the direct solution of the classical Hamilto-
cross-section. for the MD simulation of nanometric cutting nian equations of motion. The motions of the peripheral
of a single crystal of aluminum to investigate the effect of atoms are also calculated from the solution of Hamiltonian
crystallographic orientation and direction of cutting. The equations but modified by the presence of velocity reset
work material in the MD simulation is divided into three functions associated with each atom in the peripheral zone.
different zones, namely the moving zone ŽP-zone., the In this method, the Cartesian velocity components of each
peripheral zone ŽQ-zone., and the boundary zone ŽB-zone. peripheral lattice atoms is reset at periodic time intervals,
w49x. A layer of boundary and peripheral atoms are placed D t, using the following algorithm:
on all sides of the crystal except on the surface of the work 1r2
material. That is, if we consider the work material is Õanew
i s Ž1yW . Õaoldi q w 1r2 V Ž T , j . ,
64 R. Komanduri et al.r Wear 242 (2000) 60–88

where Õaoldi is the a-component Ž a s x, y, or z . of velocity tion through space rather than experience the cutting ac-
of lattice atom i resulting from the solution of the Hamil- tion. In addition, the present investigation simulates or-
tonian equations of motion and Õanewi is the reset a velocity thogonal machining, or 2-D machining, also known as
component. w is a parameter that controls the strength of machining under plane strain conditions. The fixed bound-
the reset with w s 0 corresponding to no reset and w s 1 ary conditions enable plane strain conditions or zero strain
being a complete reset. V ŽT, j . is a randomly chosen in the X-direction.
velocity from a Maxwell–Boltzmann distribution at tem-
perature T. j is a random number whose distribution is
uniform on the interval w0,1x that controls the random 3.2. MD simulation conditions
selection. This procedure simulates the thermostatic effect
of the bulk and guarantees that the equilibrium temperature MD simulations of nanometric cutting were conducted
will approach the desired value, which is 293 K in these on single crystals of aluminum using a Digital a-worksta-
calculations. More details relating to this thermostating tion ŽModel 500. with a clock speed of 433 MHz to study
method are given in Chandrasekaran et al. w44x and in the the effect of crystallographic parameters, namely crystal
original paper by Riley et al. w49x. The boundary atoms are orientation and direction of cutting. Machining studies
fixed in position and serve to reduce the edge effects and were conducted at a cutting speed of 500 mrs and differ-
maintain proper symmetry of the lattice. ent depths of cut Ž0.81–1.62 nm. along specific combina-
Generally, there are two types of boundary conditions tions of the crystal orientation Ž111., Ž110., and Ž001.4
used, namely, periodic and fixed, in MD simulation stud- and cutting direction ²w110x, w211x, w100x:. The specific
ies. In this investigation fixed boundary conditions are combinations for Ž111. orientation are the cutting direc-
used. Periodic boundary conditions are employed when a tions w110x and w211x, for Ž110. orientation are the cutting
simulation seeks to investigate the behavior of an isolated directions w110x and w001x, and for Ž001. orientation are the
crystal system, to avoid spurious edge effects and thereby cutting directions w110x and w100x, respectively. In addition,
simulate the behavior of a much larger crystal system. tools with different positive rake angles Ž08, 108, and 408.
Typical examples include studies of chemical reactions were used to investigate their effect on the mechanics of
under matrix isolation conditions. However, in the present nanometric cutting. For convenience, an infinitely hard
application, the system is not isolated. The work material Žtungsten. tool was used in these simulations as tool wear
is rigidly clamped in a vise against the external forces is hardly a problem when machining pure aluminum.
produced by the cutting tool. Proper modeling of these Table 2 gives the computational parameters, details of
external forces requires that periodic boundary conditions the work material and tool dimensions, width of cut, depth
be replaced by fixed boundary atoms. For example, a of cut, length of cut, and tool geometry used in the
model employing periodic boundary conditions but with- simulations. Fig. 2 shows the various crystal orientations
out fixed boundary atoms would simply undergo transla- and the corresponding cutting directions for cubic crystals

Table 2
Computational parameters used in the MD simulation of nanometric cutting of aluminum single crystal
Configuration 3-D cutting
Potential used Žpotential parameters Morse potential ŽD L s 0.2703 eV, a L s 1.1646rA,
for the workmaterial. reL s 3.253 A and rc s 5.624 A.
Workmaterial dimension 4a = 30a = 25a, a-lattice constant


Number of atoms in the workmaterial Crystal set-up No. of atoms


based on crystal set-up Ž111.w110x 4852


Ž111.w211x 4687


Ž110.w110x 4046


Ž110.w001x 4658


Ž001.w110x 4728
Ž001.w100x 3960
Tool dimension 4a = 12a = 20a, a-lattice constant 506 total tool atoms
Tool material Infinitely hard
Tool edge radius Sharp edge
Tool rake angle 08, 108, and 408
Tool Clearance angle 58
Depth of cut 0.81 y 1.62 nm
Width of cut 1.62 nm and 3.24 nm
Cutting speed 500 mrsec
Bulk temperature 293 K
R. Komanduri et al.r Wear 242 (2000) 60–88 65

study nanocutting at tool speeds in the experimental range


using Monte Carlo simulation techniques. At present, we
are developing computer codes to implement such studies.
We anticipate that comparison of Monte Carlo results with
MD simulations will provide a clear picture of the arti-
facts, if any, that are generated by employing high cutting
speeds.
The potential used in the simulations is a pairwise sum
of Morse potentials between the lattice atoms of the work
material plus a second summation of pairwise Morse po-
tentials between the atoms of the tool and those in the
work material, i.e.:
NW NW N W NT
Vtotal s V T s ÝÝ VML Ž ri j . q Ý Ý VMTL Ž ri j . ,
is1 j)i is1 js1

where ri j is the i–j interparticle distance and:


V ML Ž ri j . s D L exp  y2 a L Ž ri j y reL . 4

y2exp  ya L Ž ri j y reL . 4 for ri j F rc ,


V ML Ž ri j . s 0 for ri j ) rc ,
Fig. 2. Various crystal orientations and the corresponding cutting direc- V MTL Ž ri j . s D T exp  y2 a T Ž ri j y reT . 4
tions for cubic materials w47x.
y2exp  ya T Ž ri j y reT . 4 for ri j F rc ,
w47x that are used in this investigation. Fig. 3 shows V MTL Ž ri j . s 0 for ri j ) rc .
schematically the orientation of the work material with
respect to the direction of cutting w50x. It should be noted NW and NT are the number of lattice atoms in the work
that while the simulation size of the work material is material and tool, respectively. The potential parameters
maintained constant for all the crystal orientations and for the work material, D L , a L , and reL are adjusted to the
cutting directions ŽTable 2., the number of planes and measured sublimation enthalpy, Debye temperature, and
consequently, the total number of atoms considered in the nearest-neighbor spacings for aluminum. The three param-
simulation size depend on the crystal orientation. The eters for the tool-work material potential, D T , a T , and reT
number of atoms in the work material varies from 3960 to are obtained by standard combining rules with the parame-
4852 depending on the crystal setup. Table 2 also tabulates ters for the pure materials. The cutoff radius Ž rc . is chosen
the total number of atoms in the work material for various such that the ratio of the potential at the cutoff point to that
combinations of crystal orientation and cutting direction. at equilibrium is less than 4%. The use of such a cutoff
Consequently, the number of atoms to be cut for a particu- radius ensures that the calculations will not spend large
lar depth and width of cut will change with orientation and quantities of computational time to evaluate the forces that
cutting direction. While, the work material orientation is are near zero. It should be noted that there are no potential
varied, the tool orientation is maintained constant in all the terms connecting the atoms of the tool with the work
simulations ŽTable 2.. The total number of atoms in the material atoms in the boundary zone. Consequently, the
tool is 506.
The cutting speed used, namely 500 mrs, is very high
compared to conventional speeds Ž1–10 mrs. in UPM.
Unfortunately, it is necessary to use such high cutting
speeds to enable MD simulations to be conducted in a
reasonable period of time Ža few days per run instead of a
few weeks per run.. Currently, this is one of the drawbacks
of MD simulations. It is possible that high cutting speeds
may produce spurious effects. It is also possible that the
properties that we are investigating are insensitive to tool
speed so that no aphysical effects are produced. Since we
cannot execute the MD simulations at conventional cutting
speeds of 1 to 10 mrs, these theoretical methods cannot be Fig. 3. Schematic showing the orientation of the workmaterial with
used to investigate this point. However, it is possible to respect to the direction of cutting w50x.
66 R. Komanduri et al.r Wear 242 (2000) 60–88

Table 3 Ž2. MD techniques offer higher temporal and spatial


Atomic density and distance between planes for aluminum crystal Žafter resolution of the cutting process than is possible by any
Dieter, 1986.
other technique including the continuum mechanics ap-
Orientation Atomic density per Distance between
proach, finite difference methods, and FEM. Hence, any
unit area, rA2 planes, A
physical phenomena not accounted for in continuum analy-
Octahedral 1114 4r63a2 Ž0.141. ar63 Ž2.338.
sis due to size limitations can be effectively investigated
Cube 1004 2ra2 Ž0.122. ar2 Ž2.025.
Dodecahedral 1104 2r62a2 Ž0.086. ar262 Ž1.432. by MD simulation.
Ž3. MD techniques can be used effectively to simulate
the mechanics of nanometric cutting.
tool can, in effect, move through the B-zone atoms as if Ž4. MD techniques do not require the use of expensive
they are not present. The potential parameters for the work equipment, such as ultraprecision diamond turning ma-
material are given in Table 2. chine, diamond tools, and considerable experimentation to
The use of pairwise potential in the simulation merits obtain data.
some comment. Numerous trajectory investigations of Ž5. In MD simulation, since the process can be reduced
complex chemical reactions in the gas phase w51x and in to its fundamental units for analysis, MD data should
the solid state under matrix isolation conditions w52x have provide the theoretical limits of the machining process.
shown that pairwise potentials yield accurate results so Ž6. Effect of such variables as edge radius, rake angle,
long as the process under investigation does not involve and cut depth can be easily and effectively varied in MD
the scission and formation of covalent chemical bonds. If simulation. The experimental work requires expensive sin-
the machining process involves covalently bonded materi- gle crystal diamond tools, careful characterization of the
als such as silicon or diamond, more complex potential tool, and considerable testing on an ultraprecision machine
forms must be employed. Other potential forms can, of tool, which is both time consuming and costly. In MD
course, be used for any system. One such potential is the simulations, one can define these variables and conduct
embedded-atom potential. Further MD studies will employ simulations accordingly.
this formulation for the potential. When these calculations Ž7. It is easy to vary the properties of the work materi-
are complete, a comparison of the present results obtained als and cutting tools in MD simulations. For example, to
with pairwise potential with those given by an embedded- study the effect of relative hardness of the tool with
atom functional form will permit an accurate assessment of respect to work material, the hardness values can be
the sensitivity of the computed dynamics to the pairwise defined simply in the MD simulation but to obtain tools
assumption. with these specified hardness may be difficult, if not
Table 3 gives the atomic density and the distance impossible, in some cases.
between the planes with respect to the orientation of the Ž8. Contrary to general perception that MD simulations
aluminum crystal w1x. The atomic density for aluminum, an can only be done on perfect materials and cannot address
FCC material, is maximum for the octahedral Ž111. plane real materials, it can be shown that defect structures, such
and minimum for the dodecahedral Ž110. plane. Note that as grain boundaries, voids, second phase particles, disloca-
the planes of greatest atomic density Žatoms per unit area. tions, etc., can be accurately modeled though on a limited
also are the most widely spaced planes. Both the atomic scale and under idealistic conditions.
density and the distance between the atoms in a given Ž9. Tool-work interactions that cannot be studied by
direction of cutting are important considerations, for slip FEM techniques, can be addressed in MD simulation.
will occur on the densest planes and the shortest directions. Thus, tool wear, plastic deformation of the tool, and
Thus, a combination of both these factors affect the nature subsurface deformation of the work material can all be
of plastic deformation of this material in nanometric cut- studied for some simple systems. For example, tool wear
ting. in machining iron with a diamond tool or machining of
silicon with a diamond tool can be investigated using the
actual potential of these materials and suitable interaction
4. Advantages and current limitations of MD simula- potentials.
tion of nanometric cutting Ž10. Effect of crystal orientation of the work materialr
tool on the mechanics of chip formation in machining can
4.1. AdÕantages be investigated systematically. Similar experimental work
would be expensive due to very high cost of single crystal
Ž1. Unlike in FEM, nodes and distance between nodes materials and specimen preparation time.
in MD simulations are not selected arbitrarily but are
chosen on the basis of more fundamental units of the 4.2. Current limitations
material, namely, centers of the atoms as the nodes and
interatomic distances or lattice constant as the distance Ž1. Number of atoms considered in the work material is
between nodes. rather small Žranging from a few hundred to a few thou-
R. Komanduri et al.r Wear 242 (2000) 60–88 67

sand atoms. and a few hundred atoms for the tool. Perhaps incorporating statistical mechanical approaches, such as
a million atoms would be an upper limit for the work Monte Carlo simulation.
material at this stage and this would take considerable time Ž6. Simulation time is long Žcan be several hundred
for simulation. This limitation may still apply because of hours for each simulation.. Fortunately, with the availabil-
the computational time required, if one were to reduce the ity of reasonably low priced workstations with high com-
cutting speeds close to conventional Ž; 2 mrs.. However, putational speeds and considerable memory, simulations
it may be possible to simulate larger number by combining are done without the need for human involvement during
MD simulation with statistical mechanical approaches, such most of this period.
as Monte Carlo simulation. Some limitations of the num- Ž7. MD simulation is applied thus far mostly to quasi
ber of atoms can be overcome using techniques, such as 2-D Žmostly planar. cutting. This is not a real limitation
Length Restricted MD simulation. but a practice by some researchers. A quasi 3-D cutting
Ž2. Applied mainly to nanometric cutting, i.e., cut depth Žplane strain. should be conducted even if that involves
in the nanometer range and is not in the general range of more processing time as it is more representative of the
machining. This, however, is not a limitation especially real process. It may be noted that much of FEM analysis of
when investigating the mechanics of nanometric cutting. cutting is 2-D cutting and much of the mechanics of the
Experimental results in conventional machining can be cutting process developed thus far is 2-D.
compared to the results in nanometric cutting with appro- Ž8. In most of the MD simulation studies reported thus
priate consideration for the specimen size, or the size far, the tool is generally considered infinitely hard; hence,
effect. tool–work interactions, tool wear, tool deformation are
Ž3. Cutting speeds used are extremely high Ž100–500 neglected. However, this limitation can be easily relaxed
mrs. due to long processing times. It is without doubt that when appropriate interface parameters for the potential
the cutting speeds used in MD simulation are too high and function are developed.
somewhat unrealistic. This is a natural criticism of the MD
simulation of machining studies. However, with the ad-
vances in both hardware and software used in the MD 5. Results and discussion
simulation, it would be possible in the not too distinct
future to simulate conditions much closer to the actual 5.1. Effect of crystal orientation and direction of cutting on
cutting conditions. It may be noted that even at these nanometric cutting of single crystal aluminum
speeds, it is possible to simulate certain aspects of nano-
metric cutting realistically, which is not that easy by other In the following, the nature of deformation and the
experimental techniques andror requires extensive experi- variation of cutting forces and specific energy in nanomet-
mentation and analysis ric cutting will be presented.
Ž4. Currently, MD simulations are applied to only sim-
ple systems, such as single elements for the work material 5.1.1. On the nature of deformation in cutting
and the tool. This will be the case till potential energy Figs. 4Ža. – Žd. –9Ža. – Žd. are MD simulation plots of
surfaces are developed for other materials systems of nanometric cutting showing various stages for specific
interest, which is no trivial matter. There are, however, combinations of workpiece orientation, namely Ž111.,
considerable number of applications of nanometric cutting Ž110., and Ž001., and cutting direction, namely w110x,
of nonferrous work materials, such as Cu, Al and semicon- w211x, and w100x with a 108 rake tool. For example, for
ductor materials, such as Si and Ge with single crystal Ž111. orientation, the cutting directions used are w110x and
diamond tool. Similarly, wear of diamond in machining w211x; for Ž110. orientation they are w110x and w001x; and
iron can be simulated. Also, one can conduct MD simula- for Ž001. orientation they are w110x and w100x. It may be
tion of work materials, which are alloys whose compo- noted that the discussion of results presented in the follow-
nents are mutually soluble in both liquid and solid states, ing is based not only on the MD simulation plots of the
such as nickel–copper alloys. The distribution of the atoms various stages of cutting wFigs. 4Ža. – Žd. –9Ža. – Žd.x but also
in the work material will be proportional to the atomic on the detailed study of the animation of the nanometric
weight percent of the respective elements. cutting process under various conditions. The animations
Ž5. MD simulation is applied thus far for mostly pure were accomplished using a special program developed at
materials with no defects such as voids, grain boundaries, Oklahoma State University. It will be shown that the
dislocations, second phase particles, etc. Various aspects of nature of deformation ahead of the tool, the generation and
these are addressed recently and it is possible to simulate propagation of the dislocations, and the subsurface defor-
the presence of grain boundaries, voids, presence of sec- mation of the work material all depend very much on the
ond phase particles to a certain extent under idealistic orientation and the direction of cutting. Even the mecha-
conditions. Thus, defect structure in the work material can nism of chip formation will be shown to vary significantly
be taken into account in MD simulation. The density, size, depending on the crystal orientation and direction of cut-
and shape of the defects may have to be addressed by ting. It will also be shown that while the chip formation
68 R. Komanduri et al.r Wear 242 (2000) 60–88

Fig. 4. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x direction
and Ž111. crystal orientation. Rake angle 108.

mechanism is similar to the conventional shear ahead of minimal deformation in the vicinity of the cutting tool
the tool in the case of a polycrystalline material for the ŽFig. 4Ža.. as in conventional cutting at large depths of cut.
work material-cutting direction combination of Ž001. w100x, Apart from this, the uncut material in the workpiece seems
they are distinctly different in other combinations. It will to be little affected by the motion of the tool. The chip
also be shown that the elastic recovery along the machined formation appears to be predominantly due to compression
surface of the work material is different for different ahead of the tool combined with shear along the cutting
combinations of crystallographic orientations and cutting direction ŽFig. 4Žb. and Žc... As cutting progresses, the
directions thereby affecting the surface finish and integrity deformation ahead of the tool in the depth of cut region as
of the finished surface. Some of the elastic recovery well as below this in the workpiece leading to subsurface
relaxes the atoms in the crystal near the machined surface deformation in the machined surface Žunder the clearance
to their nearly original state thereby reducing the subsur- surface of the tool. ŽFig. 4Žb.. can be seen. It may be noted
face deformation. that Ž111. w110x combination involves the densest plane
Fig. 4Ža. – Žd. shows various stages of MD simulation of with the cutting performed along the close packed direc-
nanometric cutting of a single crystal aluminum along tion Žsee Table 3.. As a result, the material ahead of the
w110x cutting direction and Ž111. crystal orientation. Ini- tool is compressed significantly with cutting. As cutting
tially, the material is deformed ahead of the tool with progresses, not only the material ahead of the tool is
R. Komanduri et al.r Wear 242 (2000) 60–88 69

Fig. 5. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w211x direction
and Ž111. crystal orientation. Rake angle 108.

deformed but several layers below the uncut material Ž111. crystal orientation. Fig. 5Ža. shows the generation
ahead of the tool tip is affected ŽFig. 4Žb. and Žc... As the and propagation of dislocations into the work material at
tool passes this region, partial elastic recovery of the an angle of ; 608 to the cutting direction Žanticlockwise..
deformed subsurface region takes place. From Fig. 4Žd., It may be noted that the direction of dislocation motion in
which is the final stage of the cutting process, the amount this case is different to the dislocation motion observed in
of subsurface deformation appears to be nearly equal to the Fig. 4Ža. – Žd.. Though the crystal orientation is the same in
depth of cut. However, as the tool progresses further, some both cases, the cutting directions are different. Incidentally
of the subsurface deformation underneath and ahead of the for a cube system, the angle between the slip plane Ž111.
tool tip is released, resulting in the subsurface deformation and the slip direction w211x is 61852X w53x. The material in
of only about half the depth of cut was in Fig. 4Žc.x. Also, the shear zone also seems to be deforming at an angle of
much of the remaining work material is undisturbed, as no ; 608, as though a mirror image of the dislocations propa-
disorder in the crystal can be seen. The chip appears to be gating into the work material ŽFig. 5Žb. and Žc... As
thick and shorter due to predominant compression ahead of cutting progresses, multiple dislocations are observed to
the tool together with shear along the cutting direction. travel through the entire depth of the work material ŽFig.
Fig. 5Ža. – Žd. shows various stages of MD simulation of 5Žb... Fig. 5Žd. shows the final stage of the cutting process
nanometric cutting when cutting along w211x direction and with minimal subsurface deformation except for the dislo-
70 R. Komanduri et al.r Wear 242 (2000) 60–88

Fig. 6. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x direction
and Ž110. crystal orientation. Rake angle 108.

cations. However, the elastic recovery enables the move- cutting direction in the region of depth of cut of the work
ment of many of the dislocations to the free surface material as well as perpendicular to the cutting direction
Žmachined surface. resulting in minimal subsurface defor- into the work material underneath the tool. Also seen in
mation. The machined surface indicates better finish and Fig. 6Ža. as well as from the animation studies are the
integrity than that observed with Ž111. w110x combination planes within the depth of cut, which glide past one
ŽFig. 4Žd... Also, Fig. 5Žd. shows lesser amount of plastic another along the cutting direction like a deck of cards
deformation ahead of the tool in comparison to Fig. 4Žd.. sliding over one other. Consequently, there is very little
This is due to the differences in the dislocation motion as subsurface deformation underneath the tool initially. It
it leads to variations in the nature of plastic deformation. may be noted that slip of the layers within the uncut depth
This, consequently, leads to the variation in forces, force along the cutting direction Žw110x. was in Fig. 6Ža.x was not
ratio, specific energy, and subsurface deformation, as will observed with the Ž111. orientation ŽFig. 4Ža.. even though
be discussed shortly. the cutting direction was the same. In Fig. 4Žb., subsurface
Fig. 6Ža. – Žd. shows the MD simulation of nanometric deformation below the depth of cut in the work material
cutting at various stages when cutting along the w110x was observed in the initial stages itself. In contrast, Fig.
direction with Ž110. crystal orientation. It can be seen from 6Žb. shows no such deformed layer below the depth of cut
Fig. 6Ža. – Žc. that dislocations are generated parallel to the though the cutting direction is the same in both cases. Fig.
R. Komanduri et al.r Wear 242 (2000) 60–88 71

Fig. 7. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w001x direction
and Ž110. crystal orientation. Rake angle 108.

6Žc. shows the generation and propagation of additional in Fig. 6Žc. under the tool tip are no longer present in the
dislocations perpendicular Žinto the work material. to the work material although new dislocations are generated
cutting direction as a result of dislocations generated along instead. It was found from the animation of the MD
the cutting direction. It is also possible that these disloca- simulation process that many of the dislocations generated
tions may have been generated due to the boundary atoms perpendicular to the cutting direction terminated through
located farther from the tool in the cutting direction, which the free surface due to elastic recovery of the work mate-
can obstruct the dislocations propagating parallel to the rial after cutting. The initiation of the perpendicular dislo-
cutting direction. However, it will be shown ŽFig. 8Ža. – Žd.. cations resulted in minor disorder of the atoms near the
that in the case of Ž001. w110x combination, where the surface towards the end of the cutting process. However,
dislocations propagate parallel to the cutting direction some of these dislocations were observed to travel through
wsimilar to Fig. 6Ža. – Žd.x, no such dislocations propagating the entire depth of the work material. Those that could not
perpendicular to the cutting direction was observed. It, escape completely through the machined surface due to
therefore, appears probable that the dislocations propagat- elastic recovery were found to introduce minor lattice
ing perpendicular to the cutting direction in this case may disorders in the machined surface.
have been generated due to the orientation effect. Fig. 6Žd. Fig. 7Ža. – Žd. shows the MD simulation of nanometric
shows the final stage of the cutting process where the cutting at various stages when cutting along w001x direction
perpendicular dislocations into the work material observed with Ž110. crystal orientation. In contrast to the other
72 R. Komanduri et al.r Wear 242 (2000) 60–88

Fig. 8. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x direction
and Ž001. crystal orientation. Rake angle 108.

orientations and cutting directions presented earlier, the ŽFigs. 4–6Ža. – Žd.., the deformation was more of shear
dislocations are seen generated perpendicular to the cutting than compression. As cutting proceeds ŽFig. 7Žb. – Žd.., the
direction ahead of the tool and propagated both below and subsurface deformation is observed to be rather low. The
above the work material ŽFig. 7Žb... This mode of defor- dislocations Žor the subsurface deformation. observed in
mation is somewhat peculiar to single crystal material as Fig. 7Žb. or Žc. are no longer seen in Fig. 7Žc. or Žd.. From
no such effect was observed in conventional machining a study of the animation of the MD simulation process, it
with polycrystalline materials. Hence, the plastic deforma- was observed that the elastic recovery was rather rapid for
tion around the tool is constrained to this region only. As this combination. Though, a large number of dislocations
the dislocations are generated perpendicular to the cutting were observed to propagate into the work material, many
direction, the deformation ahead of the tool tip is accompa- of them escaped through the machined surface. Rest of the
nied by shear in this direction resulting in very high values dislocations deep in the work material rearranged them-
of the shear angle. This, however, was not the case with selves by the end of the process leaving behind practically
the other combinations of crystal orientations and cutting no disorder in the work material. Fig. 7Žd. shows the chip
directions discussed earlier. Though there was a difference to be thin and long with minimal subsurface deformation
in the amount of plastic deformation ahead of the tool except for the atoms at or near the surface. The chip
R. Komanduri et al.r Wear 242 (2000) 60–88 73

Fig. 9. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w100x direction
and Ž001. crystal orientation. Rake angle 108.

thickness was observed to be minimum for this particular In the case of Ž111. orientation ŽFig. 4Žb.., subsurface
combination in comparison to the combinations discussed deformation was observed in the form of lattice disorder.
earlier. In the case of Ž110. orientation ŽFig. 6Žb.. dislocations
Fig. 8Ža. – Žd. shows MD simulation plots of the various propagating into the work material perpendicular to the
stages of the nanometric cutting process along w110x direc- cutting direction were observed. However, with Ž001. ori-
tion with Ž001. crystal orientation. Similar to the observa- entation, dislocations move along the cutting direction in
tions made earlier with Ž110. w110x ŽFig. 6Ža. – Žd.. combi- the work material as well as in the depth of cut region. The
nation, dislocations propagating along the cutting direction amount of subsurface deformation increases as the cutting
are seen from Fig. 8Ža.. Fig. 8Žb. shows the stage after the proceeds along w110x direction. However, the dislocations
tool has moved a few nanometers into the work material in the subsurface continue to move along the cutting
showing extensive slip along the cutting direction. Also, direction ŽFig. 8Žc. and Žd... Rest of the work material
dislocations moving along the cutting direction in the work seems to be undisturbed by the cutting process as can be
material under the depth of cut region can be seen. How- observed from the final stage of cutting ŽFig. 8Žd... The
ever, this was not the case with the other two orientations subsurface disorder seen in Fig. 8Žb. and Žc. can still be
wŽ111. and Ž110.x machined along w110x cutting direction. seen in Fig. 8Žd. but to a lesser extent. From this, the
74 R. Komanduri et al.r Wear 242 (2000) 60–88

amount of elastic recovery seems to be very low in the generated along the cutting direction. Consequently, there
case of Ž001. orientation. The amount of minor disorder in is very little subsurface deformation. However, the sec-
the atoms at and near the subsurface is estimated to be ondary dislocations created perpendicular into the work
about the depth of cut. material do result in some degree of subsurface deforma-
Fig. 9Ža. – Žd. shows MD simulation plots of the various tion. In contrast in Ž110. crystal orientation but in the w001x
stages of the nanometric cutting process along w100x direc- cutting direction, the dislocations are generated perpendic-
tion with Ž001. crystal orientation. Fig. 9Ža. shows plastic ular to the cutting direction. Consequently, the deformation
deformation ahead of the tool and dislocations propagating ahead of the tool is accompanied by shear perpendicular to
into the work material at an angle of ; 458 to the cutting the cutting direction. Also, with Ž110. w001x combination, a
direction. Fig. 9Žb. shows multiple dislocations generated large amount of elastic recovery was observed. Conse-
from the tool tip and propagating into the work material. It quently, minimum subsurface deformation was observed
may be noted that in the case of Ž111. w211x combination, towards the end of the process suggesting this combination
the dislocations are observed to propagate at an angle of as a possibility for obtaining better finish in the finished
; 608 with respect to the cutting direction ŽFig. 5Žb... Fig. component. Table 4 summarizes the mode of dislocation
9Žc. shows an intermediate stage where the dislocations generation and propagation for various crystallographic
have propagated deep into the work material. Also, seen orientations and cutting directions.
from the figure is the elastic recovery of the surface atoms This variation in the dislocation generation and propa-
as the tool moves forward. Fig. 9Žd. shows the final stage gation can also be explained in terms of the angle between
of the cutting process where the dislocations propagate the cutting direction and the family of slip direction ²110:
into the entire depth of the work material. Disorder in the of an FCC crystal. For example, the angle between w100x
lattice arrangement of the crystal Žor subsurface deforma- cutting direction and the slip direction corresponds to
tion. immediately beneath the machined surface was ob- either 458 or 908 w53x. From the simulations, the disloca-
served to be limited to a fraction of the depth of cut. Even tions in the case of Ž001. w100x combination seem to
the dislocations generated in the work material may termi- propagate at ; 458 to the cutting direction ŽFig. 9Ža. – Žd...
nate into the machined surface as cutting progresses. In the case of Ž110. w001x combination, the dislocations
Closer examination of Figs. 4Ža. – Žd. –9Ža. – Žd. as well seem to propagate perpendicular Ž908. to the cutting direc-
as the animations of the MD simulation of nanometric tion ŽFig. 7Ža. – Žd... In both these combinations, the cut-
cutting of aluminum crystals of different crystal orienta- ting direction corresponds to the ²100: family. In the case
tions and directions of cutting indicates a significant differ- of cutting along w110x direction the angle between the
ence in the mechanism of shear as well as the shear angle cutting direction and the slip direction can be either 08,
for different orientations. For example, Fig. 9Žd. shows 608, or 908 w53x. Consequently, cutting along this particular
that the shear plane orientation is much closer to conven- direction produced dislocations either parallel to ŽFig.
tional Ž- 458.. However, in the case of Ž001. w110x combi- 8Ža. – Žd.. or both parallel and perpendicular to ŽFig. 6Ža. –
nation, dislocations propagating parallel to the cutting Žd.. the cutting direction. Finally, the angle between w211x
direction both in the depth of cut region as well as beneath cutting direction and the slip direction can be either 308,
the depth of cut region in the work material were observed. 548, 738, or 908 w53x. In the current study, the dislocations
In the case of Ž111. orientation and in the w211x direction, in the case of Ž111. w211x combination were observed to
the dislocations are generated at ; 608 to the cutting propagate at an angle of 608 to the cutting direction. Based
direction. The material in the shear zone also seems to be on the current study, it seems possible to predict the
deforming at an angle of ; 608, as a mirror image of the direction of primary dislocation propagation based on the
dislocations propagating into the work material. The shear angle between the cutting direction and the slip direction
angle for Ž111. w110x is far below the shear angle as of the crystal.
observed with Ž111. w211x. In the case of Ž110. crystal The thickness of the chip was observed to be minimum
orientation and w110x cutting direction, the dislocations are with Ž110. w001x combination and maximum in the case of

Table 4
Mode of dislocation generation and propagation as well as standard deviation of force variation for various crystal orientations and cutting directions
No. Cutting Plane Cutting Direction Dislocation Generation and Propagation Standard Deviation of Force Variation,
Nrmm= 10 2
1 Ž111. w110x Compression and shear 1.040
2 w211x ; 608 1.093
3 Ž110. w110x Parallel and perpendicular to the cutting direction 1.248
4 w001x Perpendicular to the cutting direction 2.043
5 Ž001. w110x Parallel to the cutting direction 1.233
6 w100x ; 458 0.996
R. Komanduri et al.r Wear 242 (2000) 60–88 75

Ž111. w110x combination. The variation in the chip thick- maximum resistance to slip as the distance between two
ness and the shear angle are due to the shear plane aligning Ž110. planes is minimum and Ž110. is also the less densely
itself for different orientations to enable easy glide or easy packed plane ŽTable 3.. Consequently, referring to Table 3,
material removal and consequent lower forces Žminimum one would expect the forces to be maximum on Ž110.
energy.. The differences in the dislocation motion lead to plane, which is not in agreement with the MD simulation
variations in the nature of plastic deformation and conse- results. However, it may be observed from Fig. 10Ža. that
quently the cyclic variation of forces, specific energy, the results are consistent for different depths of cut. It can
force ratio, and subsurface deformation. also be inferred from Fig. 10Ža. that for a particular
orientation, the cutting force is least along w110x direction
5.1.2. On the nature of Õariation of the forces and energy with Ž001. plane being an exception. On Ž001. plane,
Table 5 summarizes the results Žthe magnitudes of the minimum force is experienced along w100x direction. The
cutting and thrust forces, the force ratio, the resultant thrust force ŽFig. 10Žb.. does not seem to follow any
force, and the specific energy. of the MD simulation of particular trend as the cutting force. This is expected as the
nanometric cutting on single crystal aluminum for different energy required for plastic deformation is determined by
orientations, cutting directions, and depths of cut Ž0.81– the cutting force. However, the minimum and the maxi-
1.62 nm.. An infinitely hard tool with a 108 rake angle and mum thrust forces occur for the same orientations as the
a 58 clearance angle cutting was used at a velocity of 500 cutting force.
mrs. From Fig. 10Žd., it can be seen that the resultant force
Fig. 10Ža. – Že. shows the variation of the cutting and follows the same trend as the cutting force. This suggests
thrust forces, the ratio of the thrust force to the cutting that the cutting force is the major variable affected by the
force, the resultant force, and the specific energy, respec- orientation when machining is carried out with positive
tively, for different depths of cut Ž0.81–1.62 nm., different rake tools. The thrust force seems to play a negligible
crystal orientations, and different cutting directions. Among effect on the resultant force experienced by the tool. From
all planes and directions investigated, the cutting force was Fig. 10Že., it can be seen that the specific energy tends to
minimum in nanometric cutting on Ž111. plane along w110x increase with decreasing depth of cut indicating a possible
cutting direction. This can be attributed to the fact that size effect reported by many previous researchers w45,54–
Ž111. plane is the close packed plane and w110x is the close 58x. It can be seen that the specific energy is significantly
packed direction. Consequently, they constitute the easy influenced by both the crystal orientation of the work
slip system for an FCC crystal and hence result in the material and the cutting direction.
lowest values of the forces. It may be noted that along a Fig. 11Ža. – Žc. shows the variation of the cutting force
particular cutting direction Žw110x., the cutting force is with cutting distance for three different combinations of
minimum on Ž111. plane and maximum on the Ž001. workpiece orientations and cutting directions, namely Ž111.
plane. However, according to theory, Ž110. plane offers the w110x, Ž110. w110x, and Ž001. w100x, respectively. Table 4

Table 5
Results of MD simulation study conducted on aluminum single crystals for different orientations and different directions of cutting for various depths of
cut with a 108 rake tool
Crystal Cutting Cut depth, Cutting forcer Thrust forcer Thrust forcer Resultant forcer Specific energy,
orientation direction nm unit width, unit width, Cutting force unit width, Nrmm2 = 10y5
Nrmm= 10 2 Nrmm= 10 2 Nrmm= 10 2
111 y110 0.810 1.723 1.563 0.907 2.326 0.213
1.215 2.186 1.511 0.691 2.657 0.180
1.620 3.306 1.732 0.524 3.732 0.204
y211 0.810 2.395 1.728 0.722 2.953 0.296
1.215 3.174 1.664 0.524 3.584 0.261
1.620 4.010 1.289 0.321 4.212 0.248
110 y110 0.810 1.648 1.209 0.734 2.044 0.203
1.215 2.725 1.468 0.539 3.095 0.224
1.620 3.529 2.304 0.653 4.215 0.218
001 0.810 2.143 1.091 0.510 2.405 0.265
1.215 3.661 2.520 0.688 4.444 0.301
1.620 4.454 1.005 0.226 4.566 0.275
001 y110 0.810 2.145 2.080 0.970 2.988 0.265
1.215 3.116 2.617 0.840 4.069 0.256
1.620 4.228 3.110 0.736 5.249 0.261
100 0.810 1.857 1.471 0.792 2.369 0.229
1.215 2.494 1.798 0.721 3.075 0.205
1.620 3.437 1.708 0.497 3.838 0.2122
76 R. Komanduri et al.r Wear 242 (2000) 60–88

w001x, and Ž001. w110x4 , are not shown here, they exhibited
higher dynamic force variation in comparison to the com-
binations presented here ŽTable 4.. Each plot of force
variation for a particular orientation and cutting direction
was so chosen as to yield the minimum force for that
particular orientation. For example, in the case of Ž110.
orientation, w110x cutting direction, the force exhibits a
standard deviation value of 0.0125 Nrmm in comparison
to 0.0204 Nrmm with w001x cutting direction. Conse-
quently, w110x cutting direction was selected as it yields the

Fig. 10. Ža. – Že. Variation of the cutting and thrust forces, the ratio of
thrust force to cutting force, the resultant force, and the specific energy,
respectively, for different depths of cut Ž0.81–1.62 nm., different crystal
orientations, and different cutting directions. Rake angle 108.

tabulates the standard deviation of the force values from


the mean for different combinations. The standard devia-
tion was used to define the dynamic force variation for Fig. 11. Ža. – Žc. Variation of the cutting force with cutting distance for
various combinations. Even though the other combinations three different combinations of crystal orientations and cutting directions,
studied in this investigation, namely Ž111. w211x, Ž110. namely, Ža. Ž111. w110x, Žb. Ž110. w110x, and Žc. Ž001. w100x, respectively.
R. Komanduri et al.r Wear 242 (2000) 60–88 77

minimum dynamic force variation for Ž110. orientation. tion. It can be seen that Ž110. w110x combination presents
The biased method Žn-method. was used for evaluating the the highest force variation ŽFig. 11Žb... Also, the dynamic
standard deviation of the force curves, which defines how components of the force for Ž111. w110x and Ž001. w100x
widely the force values are dispersed from the average combinations ŽFig. 11Ža. and Žc.. were found to be similar
value. The formula used to determine the standard devia- although the amplitude of force variation was lower for
tion of the force values is given in the following: Ž001. w100x combination than for Ž111. w110x combination.
This result is in agreement with the results reported by To
nS x2y Ž S x2 .
( n2
, et al. w48x.

where, x is the value of the cutting force and n is the total 5.2. Effect of rake angle on nanometric cutting of single
number of values for which all the points in the curve ŽFig. crystal aluminum
11Ža. – Žc.. are considered. The standard deviations of the
force curves were observed to be 0.0104 Nrmm for Ž111. 5.2.1. On the nature of deformation in cutting
w110x combination, 0.0125 Nrmm for Ž110. w110x combi- In this series of tests, MD simulations of nanometric
nation, and 0.00996 Nrmm for the Ž001. w100x combina- cutting were conducted on single crystal aluminum for

Fig. 12. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x
direction and Ž111. crystal orientation. Rake angle 408.
78 R. Komanduri et al.r Wear 242 (2000) 60–88

different orientations, namely Ž111., Ž110., and Ž001.4 and and 4Ža.. shows that the deformation is found to be
cutting directions, namely ²w110x, w211x, w100x: by varying predominantly in front of the tool with a 108 rake while
the tool rake angle Ži.e., 08, 108, and 408.. The depth of cut with a 408 tool it is ahead of the tool as well as below the
was maintained constant at 1.62 nm. Figs. 12–17 are the uncut material. The smaller included angle thus provides
MD simulation plots of nanometric cutting with a 408 rake more uniform distribution of stresses in the work material
tool for various orientations and cutting directions, respec- at the tool tip resulting in deformation both in the shear
tively. The MD simulation plots for the 08 rake tool are not zone in the depth of cut region and below the uncut
given here as the differences in the nature of deformation material of the workpiece. With the progress of cut, the
between 08 and 108 rake tools are not significant. subsurface deformation is found to be somewhat similar to
Fig. 12Ža. – Žd. shows MD simulation plots of the vari- a 108 rake tool although more with the latter. Comparison
ous stages of the nanometric cutting process along the of Fig. 12Žb. – Žc. with Fig. 4Žb. – Žc. suggests lower amount
w110x direction with Ž111. crystal orientation. A compari- of plastic deformation ahead of the tool with a 408 rake
son of the same combination of crystal orientation and tool than with a 108 rake tool. Further, as can be expected,
cutting direction but with different rake angles wFig. 12Ža. – the shear angle with a 408 rake tool is much higher Žand
Žd. for 408 rake with Fig. 4Ža. – Žd. for 108 rakex indicate the chip longer. ŽFig. 12Žd.. than that with a 108 rake tool
some similarities but other differences. For example, a ŽFig. 4Žd... Fig. 12Žd. shows the final stage of the cutting
comparison of the initial stages of deformation ŽFigs. 12Ža. process where the amount of subsurface deformation seems

Fig. 13. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w211x
direction and Ž111. crystal orientation. Rake angle 408.
R. Komanduri et al.r Wear 242 (2000) 60–88 79

to be somewhat lesser than that observed with a 108 rake no longer observed in Fig. 13Žc.. This indicates that elastic
tool ŽFig. 4Žd... recovery has taken place thus relieving these dislocations.
Fig. 13Ža. – Žd. shows MD simulation plots of the vari- However, no significant elastic recovery was observed in
ous stages of the nanometric cutting process along the Fig. 5Ža. – Žd.. With a 108 rake tool, the dislocations gener-
w211x direction with Ž111. crystal orientation. From Fig. ated traveled through the depth of the work material and
13Ža. and Žb., it can be noted that the dislocations propa- could not reach the surface by elastic recovery. But in the
gate at about the same angle Ži.e., ; 608, measured anti- case of a 408 rake tool, the dislocations have propagated no
clockwise with respect to the cutting direction. as observed more than a few layers beneath the tool tip. Consequently,
earlier ŽFig. 5Ža. and Žb.. with a 108 rake tool. The the dislocations escape through the machined surface as
material ahead of the tool in the shear zone was deforming the tool advances. Fig. 13Žd. shows the final stage of the
at an angle of ; 608, as though a mirror image of the cutting process. In comparison to Fig. 5Žd., Fig. 13Žd.
dislocations propagating into the work material. However, shows minimal subsurface deformation with very few dis-
the number of dislocations observed in Fig. 13Žb. is far locations. This shows that the tool geometry plays an
less than that observed in Fig. 5Žb. though the combination important role in addition to the orientation and cutting
studied is the same except for the tool rake angle. Also, the direction effects. Hence, the results observed for the same
subsurface dislocations observed in Fig. 13Ža. and Žb. are orientation and cutting direction can vary significantly, if

Fig. 14. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x
direction and Ž110. crystal orientation. Rake angle 408.
80 R. Komanduri et al.r Wear 242 (2000) 60–88

not in the nature, in the magnitude. The subsurface defor- planes at the upper end. Consequently, this leads to an
mation appears to be minimal with this tool geometry. increased amount of cross slip and hence, more disloca-
Fig. 14Ža. – Žd. shows MD simulation plots of the vari- tions were observed to propagate into the work material
ous stages of the nanometric cutting process along w110x perpendicular to the cut depth ŽFig. 14Žd... The subsurface
direction with Ž110. crystal orientation. Fig. 14Ža. shows deformation is also seen to be minimal as most of the
the dislocations moving parallel along the cutting direction dislocations generated perpendicular to the cutting direc-
similar to the one observed with a 108 rake tool ŽFig. 6Ža... tion are relieved due to elastic recovery.
However, as the cutting progresses, an increasing number Fig. 15Ža. – Žd. shows MD simulation plots of the vari-
of dislocations perpendicular to the cutting direction are ous stages of the nanometric cutting process along the
generated ŽFig. 14Žb., Žc. and Žd... This is due to the w001x direction with a Ž110. crystal orientation. Fig. 15Ža. –
differences in the chip formation process for different rake Žd., similar to Fig. 7Ža. – Žd., shows dislocations generated
tools as explained earlier walso can be seen by comparing in front of the tool. However, with a 408 rake tool ŽFig.
Figs. 14Žb. – Žc. and 6Žb. – Žc.x. With a 108 rake tool, all the 15Ža. – Žd.. the chip formation process is in such a way that
atom layers in the depth of cut were observed to glide the dislocations generated ahead of the tool moved towards
along the cutting direction ŽFig. 6Ža... In the case of a 408 the free surface rather than into the work material. The
rake tool, the planes at the lower end of the depth of cut animations of the MD simulation process showed some of
glide along the cutting direction initially followed by the the dislocations escaping into the free surface of the

Fig. 15. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w001x
direction and Ž110. crystal orientation. Rake angle 408.
R. Komanduri et al.r Wear 242 (2000) 60–88 81

Fig. 16. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w110x
direction and Ž001. crystal orientation. Rake angle 408.

unmachined part of the work material with the remaining Fig. 16Ža. – Žd. shows MD simulation plots of the vari-
carried away by the chip. Also, no dislocations propagat- ous stages of the nanometric cutting process along the
ing into the work material perpendicular to the cutting w110x direction with Ž001. crystal orientation. The disloca-
direction are seen ŽFig. 15Ža. – Žd.. in contrast to disloca- tions move parallel to the cutting direction and the planes
tions moving into the work material perpendicular to the within the depth of cut glide over each other ŽFig. 16Ža...
cutting direction with a 108 rake tool ŽFig. 7Ža. – Žd... As As the cutting progresses, more dislocations are generated
already mentioned, this mode of deformation is rather in the work material and all the dislocations propagate
unusual in machining and is peculiar to single crystal along the cutting direction ŽFig. 16Žb. and Žc... The amount
material. Fig. 15Žd. shows very low subsurface deforma- of subsurface deformation seems to be much less ŽFig.
tion in comparison to Fig. 7Žd. due to the absence of 16Žd.. than that produced by a 108 rake tool ŽFig. 8Žd...
perpendicular dislocations propagating into the work mate- The rest of the work material remains completely unaf-
rial. Most of the deformation process and the dislocations fected.
are generated within the depth of cut and are carried away Fig. 17Ža. – Žd. shows MD simulation plots of the vari-
by the chip leaving a better finish and the work material ous stages of the nanometric cutting process along w100x
with practically very few residual defects. direction with Ž001. crystal orientation. Fig. 17Ža. shows
82 R. Komanduri et al.r Wear 242 (2000) 60–88

Fig. 17. Ža. – Žd. MD simulation plots showing various stages of nanometric cutting of an aluminum single crystal workmaterial cutting along w100x
direction and Ž001. crystal orientation. Rake angle 408.

plastic deformation ahead of the tool and dislocations ŽFig. 9Ža. – Žd... As can be seen from Fig. 17Žc., disloca-
propagating into the work material. The dislocations prop- tions generating at ; 458 Žwith respect to the cutting
agate at an angle of ; 458 with respect to the cutting direction. align and propagate parallel to the shear plane as
direction. Fig. 17Žb. shows multiple dislocations generated in conventional machining with a polycrystalline work
from the tool tip and propagating into the work material. It material at normal depths of cut. However, in the case of
may be noted that in the case of Ž111. w211x combination 108 rake tool, the dislocations were observed to travel into
the dislocations are observed to propagate at an angle of the work material and ahead of the tool tip at ; 908 to the
; 608 with respect to the cutting direction wFig. 13Žb.x. shear plane Ži.e., shear plane at ; 458 relative to the
The extent of dislocation generation and the depth to cutting direction and dislocations at 458 below cutting
which they propagate into the work material are signifi- direction..
cantly lower for a 408 rake tool ŽFig. 17Žc. and Žd.. than In summary, the chip length was observed to increase
for a 108 rake tool ŽFig. 9Žd... Fig. 17Žc. shows another and the subsurface deformation to decrease for the same
case of generation of cross slip where dislocations are orientations as the rake angle is increased. Also, the num-
generated at right-angle to the first set of dislocations in ber of dislocations and the distance traveled into the work
the work material. However, this was not observed for the material were observed to decrease with increasing tool
same orientation and cutting direction for a 108 rake tool rake angle. In the case of Ž110. w001x combination, the
R. Komanduri et al.r Wear 242 (2000) 60–88 83

dislocations were observed to generate within the cut depth 408 rake tool. This result is more significant with the thrust
and escape through the unmachined free surface of the force and the force ratio ŽFig. 19Žb. and Žc.. as rake angle
work material and the rest being carried away by the chip. more or less controls the magnitude of the thrust force,
No dislocations were observed to propagate into the work increasing more rapidly with decreasing rake angle. The
material and the subsurface damage was observed to be resultant force shows much larger variation with orienta-
minimum with practically no visible disorder in the work tion for lower rake tools than for a 408 rake tool ŽFig.
material near the finished surface. Generation of cross slip 19Žd... The specific energy follows the same trend as the
was observed in the case of Ž001. w100x combination, cutting force with least variation with orientation for the
which was not observed for the similar combination with a 408 rake tool ŽFig. 19Že... This may be one of the reasons
108 rake tool. why some researchers could not find much variation in the
forces, shear angle, and DSS with variation in crystal
5.2.2. Nature of Õariation of the forces and energy orientation and cutting direction at they used higher rake
Table 6 summarizes the results Žmagnitudes of the angles.
cutting and thrust forces, the resultant force, the force
ratio, and the specific energy. of the MD simulation of
nanometric cutting on aluminum single crystals for differ- 5.3. Effect of width of the work material
ent orientations and cutting directions with different tool
rake angles Ž08, 108, and 408.. The depth of cut was In order to investigate the effect of the width of the
maintained constant at 1.62 nm. The cutting speed used work material, simulations were conducted with increased
was 500 mrs. width of the work material Ž3.24 nm.. Based on these
Fig. 18Ža. – Že. shows the variation of the cutting and the simulations, it was observed that the shear angle as well as
thrust forces, force ratio, the resultant force, and the the deformation ahead of the tool remained essentially
specific energy for different rake angles, respectively. It unchanged with increase in the width of the work material.
can be seen from Fig. 19Ža. – Žc. that the forces and the It was also observed that the orientation of the dislocations
force ratio tend to show very little variation with orienta- with the cutting direction did not change with change in
tion and cutting direction when the rake angle is high the width of the work material. A decrease in the extent of
Ž; 408.. This is partly due to overall lower forces with a dislocations and subsurface damage was observed with
408 rake compared to a 108 rake. However, as the rake increasing work material width. This in part is due to
angle decreases, the anisotropy of forces and the force stacking up of more number of layers, which made it
ratio become significant. For example, Fig. 19Ža. shows difficult to observe the dislocation propagation. Also, an
the variation in the cutting force with orientation indicating increase in the number of atoms along the width direction
higher values with a 08 and 108 rake tool compared to a increases the force required for the dislocations to propa-

Table 6
Results of MD simulation study conducted on aluminum single crystals for different crystal orientations, different cutting directions, and different rake
angles at a depth of cut of 1.62 nm
Tool rake Crystal Cutting Cutting forcer Thrust forcer Thrust forcer Resultant forcer Specific energy,
angle, orientation direction unit width, unit width, Cutting force unit width, Nrmm2 = 10y5
degrees Nrmm= 10 2 Nrmm= 10 2 Nrmm= 10 2
08 111 y110 3.324 2.840 0.854 4.372 0.205
y211 4.110 2.490 0.606 4.805 0.254
110 y110 3.643 2.762 0.758 4.487 0.225
001 4.341 2.275 0.524 4.900 0.268
001 y110 4.237 3.863 0.912 5.734 0.262
100 3.440 2.175 0.632 4.070 0.212
108 111 y110 3.306 1.732 0.524 3.732 0.204
y211 4.010 1.289 0.321 4.212 0.248
110 y110 3.529 2.304 0.653 4.215 0.218
001 4.454 1.005 0.226 4.566 0.275
001 y110 4.228 3.110 0.736 5.249 0.261
100 3.437 1.708 0.497 3.838 0.2122
408 111 y110 2.548 1.197 0.470 2.815 0.157
y211 3.204 1.136 0.355 3.399 0.198
110 y110 2.879 1.207 0.419 3.122 0.178
001 3.106 1.083 0.349 3.289 0.192
001 y110 3.78 1.557 0.412 4.088 0.233
100 2.867 1.575 0.549 3.271 0.177
84 R. Komanduri et al.r Wear 242 (2000) 60–88

shear angle and dislocation generation and propagation,


remained unchanged. Also, many dislocations were formed
in plane than out of plane as observed from an examina-
tion of the atoms in each plane.
Table 7 summarizes the results of MD simulation,
namely the cutting force, specific energy, and the statisti-
cal error % for the two different width of cut simulations
Ž1.62 and 3.24 nm. for three different crystal setup, namely
Ž111. w110x, Ž111. w211x, and Ž001. w100x. It can be noted
that the variation in the forces and specific energy is
- 10%. It can also be noted that amongst the three crystal
setups: the cutting force is minimum with Ž111. w110x
combination, followed by Ž001. w100x combination and
Ž111. w211x combination. This trend is in agreement with
the trends observed when the width of the work material
was 1.62 nm. Consequently, based on the animations and
the forces, it appears that while the width of the work
material may have some effect on the magnitude of the
observed phenomenon and collected data, it does not seem
to affect the basic mechanics of the process significantly.

5.4. Modes of deformation in machining single crystals of


different orientations and cutting directions

Based on the detailed examination of the chip formation


process in MD simulation of nanometric cutting of single

Fig. 18. Ža. – Že. Variation of the cutting and thrust forces, the ratio of
thrust force to cutting force, the resultant force, and the specific energy,
respectively, for different crystal orientations, different cutting directions,
and different rake angles at a depth of cut of 1.62 nm.

gate into the work material due to increased number of


atomic bonds. Consequently, it can be said that while some Fig. 19. Ža. – Žc. Schematic showing three modes of plastic deformation in
differences were observed by increasing the width of the the shear zone in nanometric cutting of single crystals in different crystal
work material, the basic process phenomenon, such as the orientations and different cutting directions.
R. Komanduri et al.r Wear 242 (2000) 60–88 85

Table 7
Estimation of the statistical error % with change in width of the work material
Variables Crystal set-up Width of cut Statistical error %
1.62 nm 3.24 nm
Cutting forcerunit width, Nrmm= 10 2 Ž111.w110x 1.723 1.854 7.6
Ž111.w211x 2.395 2.539 6.01
Ž001.w100x 1.857 2.019 8.72
Specific energy, Nrmm2 = 10y5 Ž111.w110x 0.213 0.229 7.51
Ž111.w211x 0.296 0.313 5.74
Ž001.w100x 0.229 0.249 8.73

crystal aluminum, the nature of deformation is found to imental techniques while animation of the MD simulation
vary significantly with the crystal orientation and cutting of nanometric cutting permits such unusual observations.
direction. This is because in the case of single crystal
materials, slip occurs on a preferred plane for a given
6. Conclusions
orientation and direction of cutting.
Fig. 19Ža. – Žc. shows schematically three modes of Molecular Dynamics ŽMD. simulation of nanometric
plastic deformation in the shear zone in nanometric cutting cutting on single crystal aluminum were conducted in
based on the current investigation. Fig. 19Ža. shows the specific combination of crystal orientations, namely Ž111.,
case of chip formation process and the shear plane orienta- Ž110., and Ž001.4 , and cutting directions, namely ²w110x,
tion similar to the one commonly experienced in con- w211x, and w100x: with cutting tools of different positive
ventional machining of polycrystalline materials at conven- rake angles Ž08, 108, and 408. and different depths of cut
tional depths of cut. In this investigation, the chip Ž0.81–1.62 nm., at two different widths of cut Ž1. 62 and
formation process and the shear angle orientation, in the 3.24 nm., and at a cutting speed of 500 mrs to investigate
case of Ž001. orientation and w100x direction of cutting was the degree of anisotropy of this material. Specific combina-
found to be very similar to the conventional shear ahead of tion used are for Ž111. orientation, the cutting directions
the tool material as shown in Fig. 19Ža.. Thus, it is are w110x and w211x; for Ž110. orientation, the cutting
recommended that this combination be used for simulating directions are w110x and w001x; and for Ž001. orientation,
machining if only one orientation has to be used. However, the cutting directions are w110x and w100x. Specific conclu-
in certain other combinations of crystal orientation and sions that may be arrived based on this study are given in
cutting direction, the shear angles was found to be much the following.
higher than 458. For example, in the case of Ž110. orienta- Ž1. When the aluminum crystal was oriented in Ž111.
tion and w001x direction, the dislocations are generated at plane and cut in the w110x direction, the plastic deformation
right angle to the direction of cutting, i.e., material is ahead of the tool was found to be predominantly compres-
deformed perpendicular to the cutting direction. Conse- sion along with shear in the cutting direction. As cutting
quently, the shear angle is much higher than 608. This progresses, not only is the material ahead of the tool in the
mode of dislocation generation and plastic deformation in depth of cut region deformed but also several layers below
the shear zone, which is rather unique and not known in the depth of cut region ahead of the tool tip. As the tool
the conventional machining with polycrystalline work ma- progresses, some of the dislocations causing subsurface
terials, is shown schematically in Fig. 19Žb.. In contrast, deformation underneath as well as ahead of the tool tip are
when machining in Ž001. orientation and w110x direction, released, resulting in the subsurface deformation of about
the dislocations are generated parallel to the cutting direc- half the depth of cut.
tion. Consequently, the material is deformed parallel to the Ž2. When the work material was oriented in the Ž001.
cutting direction and the shear angle is much lower than plane and cut in w110x direction, the dislocations were
458. This mode of dislocation generation and plastic defor- found to generate parallel to the cutting direction. In
mation in the shear zone is shown schematically in Fig. contrast, when the crystal was oriented in Ž110. plane but
19Žc.. cut in w001x direction, the dislocations were generated
Such variations in the mode of deformation were not normal to the cutting direction. When the crystal is ori-
reported earlier as most of the machining tests were con- ented in Ž110. plane and cut in w110x direction, dislocations
ducted on polycrystalline aluminum. Even though some were found to generate both along and normal to the
researchers w10,14x reported shear angle ) 458 and as high cutting direction. Although movement of dislocations to
as 608, no explanation was offered for this unusual behav- the machined surface takes place due to elastic recovery,
ior of single crystal materials. This is because it is not easy latter orientation, namely Ž110. plane cut in w110x direction
to observe the cutting process continuously by other exper- seems to provide minimal subsurface deformation.
86 R. Komanduri et al.r Wear 242 (2000) 60–88

Ž3. In the case of Ž001. orientation and w100x cutting right angle to the direction of cutting. Material is deformed
direction, extensive dislocation motion at ; 458 to the perpendicular to the cutting direction. The shear angle is
cutting direction was observed while in the case of Ž111. ) 608. Similarly, when machining in the Ž001. orientation
orientation and w211x cutting direction, the dislocation mo- and w110x direction, the dislocations are generated parallel
tion at ; 608 to the cutting direction was observed. to the cutting direction. This is because in the case of
Ž4. The cutting forces were found to vary cyclically single crystal materials, slip occurs on a specified plane for
with the orientation of the crystal and the direction of a preferred orientation and direction of cutting. Such varia-
cutting. Minimum cutting forces were generated when tions in the mode of deformation were not reported earlier
machining in w110x directions on Ž111. or Ž110. orienta- as most of the machining tests were conducted on poly-
tions. This is followed by cutting in Ž001. direction on crystalline aluminum. Even though some researchers re-
w100x orientation. The cutting force was maximum when ported shear angle ) 458, no explanation was offered for
machining in w001x direction on Ž110. orientation. this unusual behavior of single crystal materials. This is
Ž5. A systematic variation of the cutting forces and because it is not possible to observe in situ by other
specific energy with crystal orientation and direction of experimental techniques, while animation of the MD simu-
cutting was observed. Variation in the magnitude of the lation of nanometric cutting permits observation of such
cutting force for a given depth of cut with different unusual phenomena.
orientations and directions of cutting was in the range of Ž10. Based on the MD simulation of nanometric cutting
25%, which is close to the anisotropy of this material in in different crystallographic planes and cutting directions,
the elastic range Ž21.9%.. Variation of specific energy was three modes of deformation were observed in the shear
found to range from 25.8% to 40%, which is higher than zone. In the first case, namely wŽ001. orientation and w100x
the anisotropy in the elastic range. The lower degree of directionx, the chip formation process and the shear plane
anisotropy in the case of aluminum, an FCC material, may orientation were similar to the one commonly experienced
be attributed to the high stacking fault energy Ž200 = 10y3 in conventional machining of polycrystalline materials at
J my2 . and low equilibrium spacing Žtwo atomic layers. conventional depths of cut. In the second case, namely
enabling easy cross slip. wŽ110. orientation and w001x directionx, the dislocations are
Ž6. Subsurface deformation was observed to vary with generated at right angle to the direction of cutting, i.e.,
different orientations of the crystal and direction of cutting. material is deformed perpendicular to the cutting direction.
It is seen to be minimum when the dislocation motion is Consequently, the shear angle is extremely high Ž- 608..
either parallel or perpendicular to the cutting direction. In In the third case, namely wŽ001. orientation and w110x
the former case, no dislocations were generated in the directionx, the dislocations are generated parallel to the
subsurface while in the latter case, the dislocations gener- cutting direction. Consequently, the material is deformed
ated into the work material are seen to move towards the parallel to the cutting direction and the shear angle is
machined surface due to elastic recovery. ) 458. Such variations in the mode of deformation were
Ž7. A decrease in the cutting force with increase in the not reported earlier as most of the machining tests were
rake angle was observed. The forces, force ratio, and the conducted on polycrystalline aluminum. In addition, it is
specific energy showed least variation with orientation for somewhat difficult to observe the nanometric cutting pro-
higher rake tools Ž408.. However, as the rake angle is cess continuously by other experimental techniques while
decreased Ži.e., 108 and 08., high anisotropy in the magni- animation of the MD simulation of nanometric cutting
tude of the forces, force ratio, and the specific energy with provides such an unique opportunity.
orientation is observed.
Ž8. When the dislocations move at an angle to the work
material, as in the case of Ž001. orientation and w100x Acknowledgements
cutting direction or in the case of Ž111. orientation and
w211x cutting direction, they propagate into the work mate- This project is sponsored by grants from the Manufac-
rial at depths several times the depth of cut. However, with turing Processes and Machines Program ŽDMI-9523551.
a rake tool of 408, the dislocation motion was limited to of the Division of the Design, Manufacture, and Industrial
only a few layers beneath the tool tip. This suggests that Innovation ŽDMII. and the Tribology and Surface Engi-
tool geometry also plays an important role in addition to neering Program ŽCMS 9414610. of the Division of Civil
the orientation and cutting direction effects. Hence, the and Mechanical Structures ŽDCMS. of the National Sci-
results observed for the same orientation and cutting direc- ence Foundation. The authors thank Drs. K. Rajurkar,
tion can vary significantly, if not in the nature, in the Delcie Durham, L. Martin-Vega and B.M. Kramer of
magnitude. DMII and Dr. J. Larsen Basse of the Tribology and
Ž9. Shear angles much higher than 458 were observed Surface Engineering Program for their interest in and
when machining is specific planes and cutting directions in support of this work. One of the authors ŽR.K.. also thanks
the machining of single crystal aluminum. For example, in the MOST Chair ŽMost Eminent Scholars Program. for
Ž110. orientation and w001x direction, the dislocations are at enabling the preparation of the manuscript. Special thanks
R. Komanduri et al.r Wear 242 (2000) 60–88 87

are due to Mr. Robert Stewart, a graduate student at OSU w24x X. Chang, D. Thompson, L.M. Raff, Hydrogen-atom migration on a
for the development of an animation program for the MD diamond Ž111. surface, J. Chem. Phys. 100 Ž1994. 1765.
w25x M. Perry, L.M. Raff, Theoretical studies of elementary chemisorp-
simulation of namometric cutting, which was used in this tion reactions on an activated diamond ledge surface, J. Phys. Chem.
investigation. 98 Ž1994. 4375.
w26x M. Perry, L.M. Raff, Theoretical studies of elementary chemisorp-
tion reactions on an activated diamondŽ111. terrace, J. Phys. Chem.
References 98 Ž1994. 8128.
w27x D.E. Kim, N.P. Suh, Molecular dynamics investigation of two
w1x G.E. Dieter, Mechanical Metallurgy, 3rd edn., McGraw-Hill, New dimensional atomic scale friction, Trans. ASME, J. Tribol. 116
York, NY, 1986. Ž1994. 225.
w2x R.W. Hertzberg, Deformation and Fracture Mechanics of Engineer- w28x W.G. Hoover, Molecular dynamics, in: Lect. Notes Phys. vol. 258
ing Materials, 4th edn., Wiley, New York, 1996. Springer-Verlag, Berlin, 1986, p. 13.
w3x J.T. Black, On the fundamental mechanism of large strain plastic w29x W.G. Hoover, C.G. Hoover, I.F. Stowers, Interface tribology by
deformation, Trans. ASME, J. Eng. Ind. 93 Ž1971. 507. nonequilibrium molecular dynamics, Fabr. Technol., Mat. Res.
w4x S. Ramalingam, J. Hazra, Dynamic shear stress-analysis of single Symp. 140 Ž1989. 119.
crystal machining studies, Trans. ASME, J. Eng. Ind. 95B Ž1973. w30x W.G. Hoover, A.J. De Groot, C.G. Hoover, I.F. Stowers, T. Kawai,
939. B.L. Holian, T. Boku, S. Ihara, J. Belak, Large scale elastic-plastic
w5x B.T. Chao, K.J. Trigger, L.B. Zylstra, Thermophysical aspects of indentation simulations via non-equilibrium molecular dynamics,
metal cutting, Trans. ASME 74 Ž1952. 1039. Phys. Rev. A 42 Ž10. Ž1990. 5844.
w6x B.T. Chao, K.J. Trigger, Controlled contact cutting tools, Trans. w31x J. Belak, I.F. Stowers, A molecular dynamics model of the orthogo-
ASME, J. Eng. Ind. 81 Ž1959. 139. nal cutting process, in: Proc. ASPE Annu. Conf.,1990, p. 76,
w7x S. Kobayashi, E.G. Thomson, Some observations of the shearing Rochester, NY.
process in metal cutting, Trans. ASME 81 Ž1959. 251. w32x J. Belak, D.A. Lucca, R. Komanduri, R.L. Rhorer, T. Moriwaki, K.
w8x J.A. Williams, N. Gane, Some observations on the flow stress of Okuda, S. Ikawa, S. Shimada, H. Tanaka, T.A. Dow, J.D. Drescher,
metals during metal cutting, Wear 42 Ž1977. 341. I.F. Stowers, Molecular dynamics simulation of the chip forming
w9x J.A. Williams, Small scale effects in metal cutting, Thin Films process in single crystal copper and comparison with experimental
Tribol. Ž1993. 289. data, Proc. ASPE Annu. Conf. Ž1991. 100.
w10x K. Ueda, K. Iwata, Chip formation mechanism in single crystal w33x J. Belak, Nanotribology: modelling atoms when surfaces collide, in:
cutting of beta-brass, Ann. CIRP 41 Ž1. Ž1980. 65. Energy & Technology Review, Lawrence Livermore Laboratory,
w11x W.B. Lee, Prediction of micromachining force variation in ultrapre- Livermore, CA, 1994, p. 13.
cision machining, Precis. Eng. 12 Ž1. Ž1990. 25. w34x J. Belak, I.F. Stowers, 1995, private communication.
w12x W.B. Lee, M. Zhou, A theoretical analysis of the effect of crystallo- w35x N. Ikawa, S. Shimada, H. Tanaka, G. Ohmori, An atomistic analysis
graphic orientation on chip formation in micromachining, Int. J. of nanometric chip removal as affected by tool–work interaction in
Mach. Tool Manuf. 33 Ž1993. 439. diamond turning, Ann. CIRP 40 Ž1. Ž1991. 551.
w13x W.B. Konig, N. Senrath, The influence of the crystallographic w36x S. Shimada, N. Ikawa, G. Ohmori, H. Tanaka, Molecular dynamics
structure of the substrate material on surface quality and cutting analysis as compared with experimental results of micromachining,
forces in micromachining, in: Progression Precision Engineering, Ann. CIRP 41 Ž1. Ž1992. 117.
Springer-Verlag, Braunschweig, Germany, 1991, p. 141. w37x S. Shimada, N. Ikawa, H. Tanaka, G. Ohmori, J. Uchikoshi, H.
w14x T. Moriwaki, N. Sugimura, K. Manabe, K. Iwata, A study on Yoshinaga, Feasibility study on ultimate accuracy in microcutting
orthogonal micromachining of single crystal copper, Trans. NAMRIr using molecular dynamics simulation, Ann. CIRP 42 Ž1. Ž1993. 91.
SME Ž1991. 177. w38x S. Shimada, N. Ikawa, H. Tanaka, J. Uchikoshi, Structure of micro-
w15x T. Moriwaki, K. Okuda, J.G. Shen, Study of ultraprecision orthogo- machined surface simulated by molecular dynamics analysis, Ann.
nal microdiamond cutting of single crystal copper, JSME Int. J., Ser. CIRP 43 Ž1. Ž1994. 51.
C 36 Ž3. Ž1993. 400. w39x S. Shimada, Molecular dynamics analysis of nanometric cutting
w16x B. Alder, T. Wainwright, Studies in molecular dynamics: I. General process, Int. J. Jpn. Soc. Precis. Eng. 29 Ž4. Ž1995. 283.
method, J. Chem. Phys. 31 Ž2. Ž1959. 459. w40x T. Inamura, H. Suzuki, N. Takezawa, Cutting experiments in a
w17x B. Alder, T. Wainwright, Studies in molecular dynamics: II. Behav- computer using atomic models of copper crystal and a diamond tool,
ior of a small number of elastic spheres, J. Chem. Phys. 33 Ž5. Int. J. Jpn. Soc. Precis. Eng. 25 Ž4. Ž1991. 259.
Ž1960. 1439. w41x T. Inamura, N. Takezawa, N. Taniguchi, Atomic-scale cutting in a
w18x B.W. Dodson, Molecular dynamics modeling of vapor-phase and computer using crystal models of copper and diamond, Ann. CIRP
very-low-energy ion-beam crystal growth processes, Solid State 41 Ž1. Ž1992. 121.
Mater. Sci. 16 Ž2. Ž1990. 115. w42x T. Inamura, N. Takezawa, Y. Kumaki, Mechanics and energy
w19x U. Landman, W. Luedtke, N. Burnham, R. Colton, Atomistic mech- dissipation in nanoscale cutting, Ann. CIRP 42 Ž1. Ž1993. 79.
anisms and dynamics of adhesion, nano-indentation, and fracture, w43x T. Inamura, N. Takezawa, Y. Kumaki, T. Sata, On a possible
Science 248 Ž17. Ž1990. 454. mechanism of shear deformation in nanoscale cutting, Ann. CIRP 43
w20x J. Peploski, D. Thompson, L.M. Raff, Molecular dynamics studies Ž1. Ž1994. 47.
of elementary surface reactions of C 2 H 2 and C 2 H in low-pressure w44x N. Chandrasekaran, A. Noori-Khajavi, L.M. Raff, R. Komanduri, A
diamond film formation, J. Chem. Phys. 96 Ž1992. 8538. new method for molecular dynamics simulation of nanometric cut-
w21x L.M. Raff, Effects of lattice morphology upon reaction dynamics in ting, Philos. Mag. B 77 Ž1. Ž1998. 7.
matrix-isolated systems, J. Chem. Phys. 97 Ž1992. 7459. w45x R. Komanduri, N. Chandrasekaran, L.M. Raff, Some aspects of
w22x X. Chang, M. Perry, J. Peploski, D. Thompson, L. Raff, Theoretical machining with negative rake tools simulating grinding: an MD
studies of hydrogen-abstraction reactions from diamond and dia- simulation approach, Philos. Mag. B 79 Ž7. Ž1999. 955.
mond-like surfaces, J. Chem. Phys. 99 Ž1993. 4748. w46x R. Komanduri, N. Chandrasekaran, L.M. Raff, Effect of tool geome-
w23x X. Chang, D. Thompson, L.M. Raff, Minimum-energy reaction try in nanometric cutting: an MD simulation approach, Wear 219
paths for elementary reactions in low pressure diamond-film forma- Ž1998. 84.
tion, J. Chem. Phys. 97 Ž1993. 10112. w47x M. Sato, T. Yamazaki, Y. Shimizu, T. Nakagawa, T. Takabayashi,
88 R. Komanduri et al.r Wear 242 (2000) 60–88

A study on the microcutting of aluminum single crystal, JSME 34 ŽEd.., Modern Methods for Multidimensional Dynamics Computa-
ŽIII. Ž1991. 531. tions in Chemistry, World Scientific Publishing, NJ, 1998, pp.
w48x S. To, W.B. Lee, C.Y. Chan, Ultraprecision diamond turning of 266–354.
aluminum single crystals, J. Mater. Process. Technol. 63 Ž1997. 157. w53x E.A. Wood, Crystl Orientation Manual, Columbia Univ. Press, New
w49x M.E. Riley, M.E. Coltrin, D.J. Diestler, A velocity reset method of York, NY, 1963.
simulating thermal motion and damping in gas–solid collisions, J. w54x K. Nakayama, K. Tamura, Size effect in metal-cutting force, Trans.
Chem. Phys. 88 Ž1988. 5943. ASME, J. Eng. Ind. 80 Ž1968. 119.
w50x T. Shirakashi, M. Yoshino, H. Kurashima, Study on cutting mecha- w55x R. Komanduri, Some aspects of machining with negative rake tools
nism of single crystal based on simple shear plane model, Int. J. Jpn. simulating grinding, Int. J. Mach. Tool Des. 11 Ž1971. 223.
Soc. Precis. Eng. 25 Ž2. Ž1991. 96. w56x Y. Furukawa, N. Moronuki, Effect of material properties on ultra
w51x L.M. Raff, D.L. Thompson, The classical trajectory approach to precise cutting processes, Ann. CIRP 37 Ž1. Ž1988. 113.
reactive scattering, in: Theory of Chemical Reaction Dynamics Vol. w57x T. Moriwaki, K. Okuda, Machinability of copper in ultraprecision
III CRC Press, Boca Raton, FL, 1985, Chapter I. microdiamond cutting, Ann. CIRP 38 Ž1. Ž1989. 115.
w52x L.M. Raff, Theoretical investigations of chemical and physical w58x D.A. Lucca, R.L. Rhorer, R. Komanduri, Energy dissipation in
processes under matrix isolation conditions, in: D.L. Thompson ultraprecision machining of copper, Ann. CIRP 40 Ž1. Ž1991. 83.

You might also like