You are on page 1of 30

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/376032729

Comprehensive study of material removal mechanism of polycrystalline


copper during ultra-precision cutting using molecular dynamics

Article in Precision Engineering · November 2023

CITATIONS READS

0 74

8 authors, including:

Xingying Zhou Tianyu Yu


University of Nottingham Harbin Institute of Technology
1 PUBLICATION 0 CITATIONS 58 PUBLICATIONS 1,436 CITATIONS

SEE PROFILE SEE PROFILE

Guangzhou Wang Ruiyang Guo


Harbin Institute of Technology Harbin Institute of Technology
8 PUBLICATIONS 19 CITATIONS 9 PUBLICATIONS 33 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Xingying Zhou on 29 November 2023.

The user has requested enhancement of the downloaded file.


Comprehensive study of material removal mechanism of
polycrystalline copper during ultra-precision cutting using
molecular dynamics

Xingying ZHOUb, Tianyu YUa,b*, Guangzhou WANGb, Ruiyang GUOb, Qi Liuc, Yazhou SUNb,

Henan LIUb , Mingjun CHENa,b*

a
State Key Laboratory of Robotics and System, Harbin Institute of Technology, Harbin, Heilongjiang, 150001, China
b
School of Mechatronics Engineering, Harbin Institute of Technology, Harbin, Heilongjiang, 150001, China
c
Department of Design, Manufacturing and Engineering Management, University of Strathclyde, Glasgow, G1 1XQ, UK
* Corresponding author: M.Chen, T.Yu. E-mail: chenmj@hit.edu.cn, tianyuyu@hit.edu.cn

Abstract

Nano-polycrystalline copper as a unique functional and structural material has been


drawn great attention. Comprehensive study on its nano-cutting mechanism is of great
important to guide the ultra-precision machining process. In this paper, nano-cutting
simulation of polycrystalline copper was studied against the cutting depth, tool
geometry, grain size and grain orientation. The effect of cutting depth on the shear angle
of the material deformation and the proportion of HCP atoms in chips were studied.
Furthermore, nano-indentation and nano-cutting of polycrystalline copper with
different grain size were simulated. In terms of crystal orientation, the difference of
orientation in the bi-crystalline copper could form chips with different shapes and
induce different strain distribution within the grain. In addition, the size and distribution
of dislocation density are studied in the nano-cutting process of different polycrystalline
copper textures. Texture {112}<111>, {236}<385>, and {124}<211> were considered.
Keywords: Polycrystalline copper, Cutting depth, Grain orientation, Grain size, Molecular
dynamics
1. Introduction
Due to its excellent thermal and conductive properties [1], copper is widely used in
optical components [2], chemical reaction vessels [3] and graphene growth substrates
[4, 5]. To enhance their performance, study was focused on nanoscale surface finishing
[6] and geometric accuracy [7]. In order to achieve these targets, the amount of material
removed per feed during the manufacturing process also needs to be precisely
controlled in the order of tens of nanometers [8, 9]. Nano-cutting is a promising
processing technology in micro/nano-machining research field [10]. Nano-cutting is
usually processed with a diamond tool [11], which has the advantages of high
machining efficiency, high machining precision and flexibility for different shapes [12,
13]. However, when the cutting depth is as low as nano level, nano-cutting shows
different machining characteristics from macro-cutting, including the size effect [14],
grain orientation [15, 16], grain boundaries [17] and dislocation [18, 19]. The ambiguity
of material removal mechanism has gradually become an obstacle limiting further
applications of nano-cutting [20]. Therefore, it is important to elucidate the material
removal mechanism of nano-cutting.
At present, study on nano-cutting mechanism through both experiment and
simulation were explored [21, 22]. The experimental method was mainly carried out by
in-situ nano-cutting experiment platform, which was equipped with nano-cutting
machine and in-situ scanning electron microscope (SEM) [23, 24]. The shear angle of
material deformation can be obtained by observing the chip morphology of nano-
cutting [25]. Furthermore, combining the cutting force during the nano-cutting, the
corresponding material removal mechanism be explained. To some extent, the
experimental method is still based on phenomenological interpretation of material
deformation behavior, which is difficult to reveal the further internal mechanism [26].
In addition, in order to explain the material removal mechanism, it is often necessary
to conduct a large number of experiments, which also leads to a high cost [27].
Therefore, many researchers use simulation to reveal the material removal mechanism
of nano-cutting [28].
Three simulation methods mainly used are first principle [29], molecular dynamics
[30, 31] and finite element methods [32]. The first principles treat elements as made up
of electrons and uncleus, and solve the model based on quantum mechanics [33]. Due
to the complexity of the modeling, the calculations tend to be time-consuming. In
addition, the model considers too many details and may not fit well with the nano-
cutting process. Finite element method is mainly based on continuum mechanics [34],
and the computational efficiency is high. However, when the cutting depth reaches
nanometer level, the material can no longer be regarded as the continuum, so it is
unreasonable to use the finite element method to simulation the nano-cutting process.
Molecular dynamics are considered to be the most suitable simulation method for nano-
cutting due to its scale [35]. It treats atoms as the most fundamental particles that make
up materials. In contrast to first principles, molecular dynamics does not subdivide
atoms further into electrons and nuclei, which in nano-cutting is a reasonable
simplification of the model and greatly saves simulation time. On the other hand,
compared with the continuum mechanics adopted by the finite element method, the
molecular dynamics method can provide higher temporal and spatial resolution [36], so
that more subtle phenomena in nano-cutting can be observed.
Molecular dynamics method has been used to study the material removal mechanism
involved in mechanical removal of copper materials. Li et al. [37] studied the plastic
deformation mechanism of nanocrystalline copper at different nano-scratching rates.
The results show that with the increase of scratching rate, the plastic deformation degree
of the material and the chip volume increases. This phenomenon leads to the increase
of scratching force and deformation temperature, and further leads to dislocation slip,
grain boundary slip, and grain twinning. Liu et al. [38] studied the formation
mechanism of grain boundary steps in micro-cutting polycrystalline copper using
molecular dynamics. It was found that the dislocations of continuous nucleation at the
tool-workpiece interface stopped and accumulated on the grain boundary, which
resulted in the formation of grain boundary steps. Guo [39] studied the surface
formation mechanism of single crystal copper with different grain orientations in nano-
cutting. The results show that the direction of cutting and Miller index of single crystal
could determine the direction of material accumulation. Although many scholars have
used MD simulation to study the cutting of copper material, the influence of grain size
and grain orientation on dislocation, stress and strain distribution in the cutting process
of polycrystalline copper are still insufficient. This paper attempts to construct a deeper
and more comprehensive study on the nano-cutting of polycrystalline copper, revealing
the effects of cutting depth, tool geometry, grain size, grain orientation and material
texture on nano-cutting.
The remainder of this article is organized as follows: Section 2 introduces the
molecular dynamics model used in this study. In Section 3, the effects of cutting depth,
cutting edge radius and rake angle on material removal in nanocrystalline cutting
process are studied. Then, the effects of grain size and grain orientation on the stress,
strain and dislocation distribution were discussed. Furthermore, grain orientation is
extended to the study of the texture of actual materials, and the influence of different
textures on the nano-cutting process is explored. Section 4 draws the conclusions of the
study.
2. Simulation methods

2.1 MD modeling of nanocrystalline copper


The Voronoi method is used to generate the model of nanocrystalline copper. The
specific modeling is carried out by ATOMSK, a molecular modeling software.
According to the lattice type and lattice constant of the elements, the software can
generate the required nanocrystal model by filling the volume box of the model with
atoms. The crystal structure type of copper is face-centered cubic (FCC) structure with
lattice constant of 3.61 A. The size of workpiece in the direction of XYZ is 20 nm×6
nm×11 nm, as shown in Fig. 1(a). In order to eliminate the influence of size effect as
much as possible, periodic boundary condition is adopted in the direction of Y. The
number of copper atoms involved in the simulation is on the level of 105. The workpiece
atoms can be divided into Newton layer, thermostatic layer and boundary layer, as
shown in Fig. 1(b). The atomic motion of Newton layer and thermostat layer obeys the
classical Newton’s laws. The boundary layer atoms mainly play the role of fixing the
workpiece in the nano-cutting process. Diamond tool was used for cutting simulation.
The crystal structure type of diamond is FCC structure and the lattice constant is 3.57
A. It can be seen from the description of the above model that the simulation system
mainly involves the interaction between three types of atoms, namely Cu-Cu, Cu-C and
C-C. These three types of interatomic interactions are described by different potential
functions. The interaction between Cu-Cu is described by embedded atom method
(EAM). The interaction between Cu-C is described by Lennard-Jones (LJ) potential
function. The interaction between C-C atoms is described by the Tersoff potential
function. Before cutting, the system was relaxed 20 ps under the NPT ( number of atoms
N, the pressure P, and the temperature T of the system remain constant) ensemble to
fully release the internal stress. The nano-cutting is performed in NVE (number of
atoms N, the total volume V, and the total energy E of the system remain constant)
ensemble, and the temperature of thermostat layer is maintained at 300 K. Molecular
dynamics simulation is implemented on LAMMPS software. The simulation results can
be analyzed by means of Atomic strain, Mises stress, CAN and DXA in OVITO
software.

Fig. 1 Establishment of nano-cutting model for polycrystalline copper. (a) Nanocrystalline copper
model based on Voronoi method; (b) Schematic diagram of nano-cutting.
2.2 Euler angle of polycrystalline copper texture
In the part 3.7, the influence of texture form on nano-cutting process was investigated.
Therefore, the Euler angle corresponding to the texture orientation should be calculated
before this, and then the nanocrystalline model can be generated according to the Euler
angle. The following takes the common plate texture as an example to introduce the
expression of texture and the corresponding Euler angle solution method. As shown in
Fig. 2(a), the rolling direction of the rolled plated is represented by RD, the transverse
direction of the rolled plated is represented by TD, and the normal direction of the rolled
surface of the rolled plate is represented by ND. The orientation of a grain in a rolled
plate can be expressed by {hkl}<uvw> if the {hkl} plane of the grain is parallel to the
rolling surface and the grain direction <uvw> is parallel to the rolling direction RD.
The Euler angle representation of grain orientation is shown in Fig. 2(b), which is
defined by rotating three coordinate systems of mutually perpendicular crystal axes.
The three perpendicular crystal axes are usually [100], [010] and [001]. It is assumed
that the initial positions of these crystal axes coincide with the direction RD, TD and
ND respectively, and reach the current position after three rotations. The first rotation
is the rotation of φ1 angle in the direction of [001]. The second rotation is the rotation
of φ angle in the direction of [100] after rotation. The third rotation is the rotation of φ2
angle in the direction of [001] after two rotations. The combination of these angles (φ1,
φ, φ2) is called the Euler angle of the grain orientation.

Fig. 2 A method for solving Euler angle for plates texture of polycrystalline copper. (a)
Representation of plate texture; (b) Euler angle representation of crystal orientation.

According to the coordinate transformation, the crystal orientation g after Euler angle
rotation can be expressed as Eq. (1). In the cubic crystal system, the grain orientation
can also be described by the value of the normal line of a crystal plane {hkl}, the value
of a crystal direction <uvw> on the crystal plane and the value of a crystal direction
<rst> in the reference coordinate system. The normal line of crystal plane {hkl}, the
crystal direction <uvw> and the crystal direction <rst> are perpendicular to each other.
Then, the grain orientation matrix can also be expressed as Eq. (2). It should be noted
that the values in the matrix of Eq. (2) are normalized, but the proportional relationship
between values remains unchanged. By combining Eq. (1) and Eq. (2), the values of
the three Euler angles corresponding to the texture {hkl}<uvw> can be obtained.
 cos(φ2 ) sin(φ2 ) 0 1 0 0   cos(φ1 ) sin(φ1 ) 0
g =  -sin(φ2 ) cos(φ2 ) 0   0 cos(φ) sin(φ)    -sin(φ1 ) cos(φ1 ) 0 
  
(1)
 0 0 1  0 -sin(φ) cos(φ)   0 0 1 

 u r h
g =  v s k  (2)
 w t l 

3. Results and discussion


3.1 Effect of cutting depth on the nano-cutting process
In nano-cutting, different cutting depth will lead to material accumulation of different
sizes, which will lead to the change of cutting force. When the cutting speed is 100 m/s,
the rake angle and clearance angle are 5 °, and the cutting edge radius is 1 nm, the
influence of cutting depth change on cutting force is explored. As the force in Y
direction is balanced, the cutting force in this direction during cutting is zero. The
cutting force of the other two components is related to the amount of chip material
removed. It can be seen from Fig. 3 that with the increase of cutting depth in nano-
cutting, the overall cutting force in the X direction approximately presents a linear
increase trend with a large increase amplitude. The cutting force in the X direction is
about 60 nN when the cutting depth is 1 nm, and increases to 150 nN when the cutting
depth increases to 5 nm. With the change of cutting depth, the change of cutting force
in Z direction is not obvious, and the change curve is approximately parabola. When
the cutting depth increases from 1 nm to 5 nm, the cutting force increases by about 20
nN in the Z direction.

Fig. 3 Variation of cutting forces at different cutting depths in nano-cutting.


For nanocrystalline copper nano-cutting, the cutting depth and grain size are in the
same order of magnitude. At this time, the cutting process is a trans-granular cutting
process. Fig. 4(a) shows the change curves of the three cutting force components in the
cutting process when cutting depth is 3 nm. It can be seen that when the cutting distance
reaches 11 nm, there is an obvious cutting force step in the X direction. By analyzing
the grain morphology of polycrystalline copper at the moment when the cutting distance
is 11 nm, it is found than trans-granular cutting is taking place at the moment, as shown
in Fig. 4(b). At this moment, the tool is penetrating grain 2 and starting to cut into grain
1. Therefore, when the cutting force distance is 11 nm, the generation of cutting force
step is caused by the cutting resistance of grain boundary between grain 1 and grain 2.
It can be seen from Fig. 4(a) that for the cutting force in the X direction, the cutting
resistance of grain 1 and 2 is about 50 nN.

Fig. 4 Cutting force variation curve in nano-cutting process. (a) The triaxial force variation of cutting
process when the cutting depth is 3 nm; (b) Nanocrystalline copper grain morphology at the cutting
distance of 11 nm.

Fig. 5 shows the strain distribution inside the nanocrystalline copper material at
different cutting depths. The results show that the strain mainly occurs at the grain
boundary in the non-contact area between the tool and workpiece. This is because the
grain boundary has a shielding effect on the grain interior, so the strain needs to
overcome the grain boundary resistance before reaching the grain interior. With the
increase of cutting depth, the strain reaching the grain increases, which shows that the
strain ring boundary inside the grain shrink inward. In the contact area between tool
and workpiece, when the cutting depth is 1 nm, the strain mainly occurs on both sides
of the arc of the cutting edge. And the shear angle of the material is about 45°, as shown
in Fig. 5(a). When the cutting depth is 2 nm, the strain mainly occurs in the machined
surface and shear band of the chip. The shear angle of the material decreases slightly to
44°, as shown in Fig. 5(b). When the cutting depth is increased to 3 nm, chips can form
a stable shear band of a certain width, but the change of shar angle is still not obvious,
as shown in Fig. 5(c). When the cutting depth continuous to increase, the plastic
deformation is severely hindered by the resistance generated by grain boundaries, and
the shear angle decrease significantly. As shown in Fig. 5(d), when the cutting depth is
4 nm, the shear angle drops to 35°, much lower than that when the cutting depth is 3
nm.
Fig. 5 Strain distribution at different cutting depths in nano-cutting. (a) 1 nm; (b) 2 nm; (c) 3 nm;
(d) 4 nm.

With the progress of cutting, the crystal lattice of polycrystalline copper is damaged
and a large number of dislocations such as spherical clusters and stacking dislocation
are generated. Therefore, it is of great significance to study the dislocation distribution
under different cutting depths. Because the atomic crystal structure type of the
dislocation is different from the crystal structure type of the material body, so the
common neighbor analysis can be used to analyze the dislocation distribution inside the
material. Fig. 6 shows the CNA images of polycrystalline copper with different cutting
depths. Since copper material is FCC structure, most of the identified atomic types in
the figure are FCC structure atoms. Through the analysis of atomic types in the material
pile up area, it can be found that the number of HCP atoms in the shear slip zone
increases with the increase of cutting depth. Since the type of stacking dislocation is
usually HCP structure, this indirectly indicates that the number of stacking dislocations
in the shear slip zone of material increases with the increase of cutting depth. It can also
be seen that the BCC type atoms usually form near the grain boundaries. Since the
crystal orientation of the atoms in the grain boundaries region are not clear, the crystal
type of the atom in the grain boundaries are not identified, as shown by the white atoms
in the figure. Fig. 7 shows the percentage of atoms in each type of crystal at different
cutting depths. The proportion of HCP atoms increased linearly with the increase of
cutting depth. When the cutting depth increased from 1 nm to 5 nm, the proportion of
HCP atoms increased from 2.3% to 4.1%. The ratio of atoms in BCC crystal structure
did not change much, maintaining the level of 1%. With the increase of other types of
atoms, the ratio of FCC crystal atoms in bulk polycrystalline copper materials generally
decreases.
Fig. 6 CNA image of polycrystalline copper at different cutting depth in nano-cutting. (a) 1 nm; (b)
2 nm; (c) 3 nm; (d) 4 nm.

Fig. 7 Atomic ratios of different grain structures at different cutting depths in nano-cutting.
3.2 Effect of cutting edge radius on surface material migration and strain distribution
In this part, the effects of four cutting tools with different cutting edge radius on
polycrystalline copper nano-cutting were compared under the condition of cutting depth
of 1 nm. Fig. 8 shows the influence of cutting edge radius on material accumulation on
polycrystalline copper surface during nano-cutting. Atoms at the same height as the
cutting edge are marked in red before the cutting begins, and the movement of material
atoms is tracked through the marked atoms during the cutting process. It can be seen
from the simulation results that when the cutting edge radius is 1 nm, most of the labeled
atoms are piled up to form chips except a small number of labeled atoms are squeezed
on the machined surface of the workpiece, as shown in Fig. 8(a). When the cutting edge
radius is 2 nm, a large number of labeled atoms are extruded on the machined surface,
and only a small number of atoms are migrated to the surface to form chips, which can
be seen in Fig. 8(b). When the cutting edge radius is increased to 3 nm, it can be seen
from Fig. 8(c) that the marking atoms are almost completely squeezed on the machined
surface, and the contact between the tool and the workpiece is in a ploughing state.
Therefore, in order to form stable chips in nano-cutting, the ratio of cutting edge radius
to cutting depth needs to be less than 2, which is stricter than macro-cutting. This may
because the friction between the atoms of the workpiece in nano-cutting makes it harder
to form chips than in conventional cutting. When the cutting edge radius is 4 nm, the
contact state between the tool and the workpiece is a state of extrusion and scratching.
The marked atoms are completely extruded to the machined surface, and the material
accumulation cannot be formed, as shown in Fig. 8(d).

Fig. 8 Effect of cutting edge radius on chip accumulation on polycrystalline copper surface. (a) 1
nm; (b) 2 nm; (c) 3 nm; (d) 4 nm.

Strain distribution is an important factor determining the plastic deformation of


polycrystalline copper. Fig. 9 shows the simulation results of strain distribution under
different cutting edge radius. When the cutting edge radius is 1 nm, the strain on the
workpiece surface is mainly concentrated under the tool and is in a clumpy shape, as
shown in Fig. 9(a). Strain belts appear in front of the tool and are parallel to the tool
rake surface. When the cutting edge radius is 2 nm, the strain concentration zone
gradually expands from the area below the cutting edge to the workpiece surface, and
the width of the concentration zone is 0.9 nm, as shown in Fig. 9(b). Orthogonal strain
belts appear below the tool. When the cutting edge radius is 3 nm, the width of the strain
concentration zone is reduced to 0.6 nm, and the direction of the strain belts are parallel
to the contact surface between the tool and workpiece, as shown in Fig. 9(c). When the
cutting edge radius increase further, the thickness of the strain concentration zone
decreases further, but the direction of the strain belts remains unchanged, as shown in
Fig. 9(d). The reason for the above phenomenon is that the increase of the cutting edge
radius changes the contact mode between the cutting tool and the workpiece, which
lead to the redistribution of atomic strain within the workpiece.

Fig. 9 Influence of cutting edge radius on strain distribution in tool-workpiece contact zone. (a) 1
nm; (b) 2 nm; (c) 3 nm; (d) 4 nm.

3.3 Effect of tool rake angle on chip morphology and cutting forces
Tool rake angle is a key parameter in tool geometry structure. In nano-cutting, the
rake angle of the tool affects the stacking process of material atoms through the tool
rake face, and then affects the chip shape and cutting force. Fig. 10 shows the influence
of different tool rake angle on chip shape and cutting force when the cutting depth is 2
nm and the cutting edge radius is 1 nm. The nano-cutting simulation of polycrystalline
copper was carried out under the conditions of tool rake angle of 0 °, 5 °, 10 °, 15 ° and
20 °. It can be seen from the results that the tool rake angle affects the chip crimp degree
in nano-cutting process. With the increase of the rake angle, the chip curl degree
decreases gradually. In addition, the chip thickness in the shear area of the material is
measured, and it is found that the chip thickness decreases with the increase of the tool
rake angle. By drawing the chip thickness in the shear area as a line graph, it is found
that the chip thickness decreases obviously when the cutting tool rake angle increases
to 10 °, and the chip thickness does not change significantly if the cutting tool rake
angle continues to increase.
In addition, the change of cutting force caused by tool rake angle is also studied. As
mentioned before, Fy always keep its value as zero during nano-cutting due to force
balance. Therefore, Fig. 10 shows the influence of the tool rake angle on the value of
Fx and Fz. It can be seen from the results that the influence of tool rake angle on Fx is
greater than that on Fz. As the tool rake angle increase from 0 ° to 20 °, Fx decreases
from 97 nN to 80 nN and Fz decreases from 56 nN to 50 nN. The rake angle of 10 ° is
an inflection point, and further increase of the rake angle will have little effect on the
cutting force.

Fig. 10 Influence of tool rake angle on chip shape and cutting force.

3.4 Effect of grain size of polycrystalline copper on the nanoindentation process


The influence of grain boundaries in nanomachining is not negligible because grain
boundaries prevent the propagation of dislocations and thus plastic deformation. When
the grain size is smaller, the number of grain boundaries per unit volume is more, and
the influence of grain boundaries on processing is more significant. In this section, the
effects of different average grain sizes on the nano-indentation process were explored,
the changes of crystal structure type, von Mises stress and hydrostatic pressure with
grain size were analyzed. Fig. 11 shows the modeling results of polycrystalline copper
using ATOMSK software, polycrystalline copper models with average grain sizes of 4
nm, 3 nm, 2.5 nm, 2 nm and 1 nm were established respectively. The dimension of
workpiece in X, Y and Z directions are 15 nm, 15 nm and 11 nm. Compared with the
previous sections, the workpiece in this section is remodeled because the nano-
indentation needs to maintain the same size in X and Y direction as much as possible.
The diameter of the nanometer indenter is 4 nm and compression velocity is 100 m/s.
Periodic boundary conditions are applied in the X and Y directions, and Z are applied
with non-periodic and shrink-wrapped with a minimum value.
Fig. 12 shows the nano-indentation curves of polycrystalline copper with different
grain size. Before simulation, the nano-indenter was located about 3 nm above the
workpiece surface, so when the indenter moves down 3 nm, it began to contact the
workpiece and began to generate extrusion force. When the nanometer indenter is in
contact with the workpiece, the force response of the workpiece with different grain
size is different. As can be seen from the result, the larger the grain size, the larger the
initial slope of the force-displacement curve. In addition, the smaller the grain size, the
smaller the extrusion force of the nano-indentation. This may be due to inverse Hall-
Petch relationship when grains are small enough. In recent years, research found that
the Hall-Petch relationship can be inversed when the grain size of nanostructured
material is small enough [40, 41]. The intrinsic nature of the inversed Hall-Petch
relationship is that the smaller the grain size is, the higher of the grain boundary volume
fraction is, leading to easier grain boundary sliding, thereby reducing the material
strengthen and exhibiting softening effect. The decrease in the nano-indentation force
can be further explained by the fact that below a critical grain size, the strengthening
caused by dislocation activity stops, and the activation of other mechanisms causes
softening, thereby reducing the force required for indentation.

Fig. 11 Modeling of polycrystalline copper with different grain sizes. Average grain size: (a) 4 nm;
(b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.
Fig. 12 Force-displacement curves of nano-indentation for polycrystalline copper with different
grain sizes.

In order to study the stress distribution of polycrystalline copper workpiece, it was


divided by the plane of nanometer indenter center. Fig. 13 shows the von Mises stress
distribution of polycrystalline copper workpiece with different grain sizes. It showed
that the maximum von Mises stress was located directly below or near the indenter. In
addition, for smaller grain size, the von Mises stress permeates deeper from the
extruded surface. Fig. 14 shows the results of hydrostatic pressure distribution. Atoms
with compressive stress are shown in blue and atoms with tensile stress are shown in
red. The maximum hydrostatic pressure is distributed below the indenter and shown as
compressive stress. With the decrease of average grain size, the gradient compressive
stress tends to be distributed in multiple directions. This is because the loose grain
boundary interface can absorb the compressive stress, leading to more evenly
hydrostatic pressure distribution in the subsurface of the workpiece. In addition, the
crystal structure of the material was analyzed by common neighbor analysis method, as
shown in Fig. 15. It can be seen that HCP structure atoms appear in the crystal under
the action of extrusion, which may be due to the occurrence of slip belts in grain interior.
The slip zone usually begins at the dislocation nucleation site at the grain boundary and
extends into the grain interior. In addition, with the decrease of grain size, the slip zone
began to coalesce into clusters. This is because that the smaller the grain size is, the
greater the chance of the intersection of each slip belt is.
Fig. 13 Von Mises stress distribution under different grain size of nano-indentations. Average grain
size: (a) 4 nm; (b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.

Fig. 14 Hydrostatic pressure distribution under different grain size of nano-indentations. Average
grain size: (a) 4 nm; (b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.
Fig. 15 CAN images of polycrystalline copper with different grain size in nano-indentation. Average
grain size: (a) 4 nm; (b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.

3.5 Effect of grain size on the nano-cutting process


The indentation tests of polycrystalline copper with different grain sizes showed that
with the decrease of grain size, the material strength changes in an inverse Hall-Petch
way, and the indentation force gradually decreased. In terms of stress re-distribution,
von Mises stress and hydrostatic pressure showed a general increase with the decrease
of grain size. Because the contact mode between the cutting edge and workpiece during
cutting is quite different from that of nano-indentation, so it will inevitably lead to a
change in the stress re-distribution. It is often related to the elastic recovery of a material
and affects the quality of the final machined surface. Based on the above consideration,
this section conducts nano-cutting simulations with different grain sizes to explore the
effects of grain size on the von Mises stress and hydrostatic pressure re-distribution
during the cutting process. In this section, the rake angle of the tool is 5 °, the
cuttingedge radius is 1 nm, the cutting depth is 2 nm, and the cutting speed is 100 m/s.
Fig. 16 shows the von Mises stress distribution during nano-cutting of polycrystalline
copper with different grain sizes. The red atoms in the figure show areas with high stress
values, and the green and blue atoms show areas with low stress values. It can be seen
that when the grain size is large, the maximum stress area is mainly located below the
tool tip. With the decrease of grain size, a large von Mises stress region appeared in the
chip region before the cutting tool rake face. This may be because when the grain size
is large von Mises is mainly borne by a single grain in contact with the tool tip, while
when the grain size is reduced the von Mises stress is shared by all the grain boundaries
in contact with the tool contact surface. Different from the increase of von Mises stress
with the decrease of grain size in nano-indentation, the grain size did not affect the
degree of von Mises stress expansion during nano-cutting, which may be caused by the
stress release due to the formation of chips. Fig. 17 shows the re-distribution of
hydrostatic pressure. The blue atoms represent the regions in compression, and the red
atoms represent the regions in tension. It can be seen that the compressive stress is
mainly located just below the tool tip, and the tensile stress is near the fixed surface.
Similar to nano-indentation, the gradient compressive stress also shows a tendency to
distribute in multiple direction when the grain size decreases. However, due to the
formation of chips, the stress is released, so compared with the nano-indention, the
compressive stress is less intense.

Fig. 16 Von Mises stress distribution in nano-cutting polycrystalline copper with different grain
sizes. Average grain size: (a) 4 nm; (b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.
Fig. 17 Hydrostatic pressure distribution in nano-cutting polycrystalline copper with different grain
sizes. Average grain size: (a) 4 nm; (b) 3 nm; (c) 2.5 nm; (d) 2 nm; (e) 1 nm.

In order to quantitatively investigate the effect of grain size on the roughness of the
machined surface, the atomic coordinates of the machined surface of the workpiece
with different grain size were extracted to calculate the surface roughness Sa. First, the
atoms on the machined surface are selected by OVITO post-processing analysis
software, as shown in Fig. 18(a). Fig. 18(b) shows the selected machined surface atoms,
and different colors represent atoms with different grain orientations. The three-
dimensional coordinates of the selected atoms are derived and the surface roughness of
the machined surface is calculated by the formula of surface roughness Sa. Fig. 18(c)
shows the result of the selected atoms being drawn again in the numerical analysis
software. Fig. 18(d) shows the variation of the surface roughness of the machined
surface with the grain size. It can be seen from the results that the smaller the grain size,
the greater the surface roughness. This may be due to the difference of material elastic
recovery rate on both sides of grain boundary after machining. The different recovery
rate of the material induces the formation of steps at the grain boundary, thus increasing
the surface roughness.
Fig. 18 The variation of surface roughness with grain size of processed polycrystalline copper. (a)
Selection of machined surface; (b) Machined surface morphology; (c) Coordinate extraction of
atoms from a machined surface; (d) The variation of surface roughness Sa with grain size.

3.6 Effect of grain orientation difference on strain distribution and cutting forces
Then density of atoms in different orientations in the crystal is different, which
affects the material removal in nano-cutting process. In this part, the strain distribution
and cutting force variation of three groups of bi-crystalline copper with different
orientations are studied. Fig. 19 shows the orientation layout diagram of three group of
bi-crystalline copper. The orientation of the first grain in the three bi-crystalline copper
is the same, and [100], [010] and [001] crystal directions are selected as its crystal axes.
The orientation difference of the three groups of grains is determined by the orientation
of the second grain. The three crystal axes of the second grain of the first group are
[210], [-120] and [001] crystal directions of the first grain, as shown in Fig. 19(a). The
three crystal axes of the second grain of the second group are the [110], [-110] and [001]
crystal directions of the first grain, as shown in Fig. 19(b). The three crystal axes of the
second grain of the third group are the [11-1], [112] and [1-10] orientation of the first
grain, which can be seen in Fig. 19(c). In this section, the rake angle of the tool is 5 °,
the cutting edge radius is 1 nm, the cutting depth is 1 nm, and the cutting speed is 100
m/s.
Fig. 19 Schematic diagram of orientation difference of three groups of different bi-crystalline copper.
(a) [100] [010] [001] with [210] [-120] [001]; (b) [100] [010] [001] with [110] [-110] [001]; (c) [100]
[010] [001] with [11-1] [112] [1-10].

Fig. 20 shows the strain distribution diagram of three groups of different bi-
crystalline copper in nano-cutting process. Two representative moments of just cutting
into the workpiece and just cutting across the grain boundary are selected for analysis.
The results show that the grain orientation not only affects the strain distribution but
also the chip shape. For the first group of bi-crystalline copper, when the tool just cut
into grain 2, the whole grain 2 has a large strain. In addition, grain 2 induced grain 1 to
strain in the crystal direction of [-101] and [101] through the action of grain boundary,
as shown in Fig. 20(a). When the tool cuts through the grain boundary, the chip
produced is typical of banded chip commonly seen in metal cutting, as shown in Fig.
20(a-1). In the second group of bi-crystalline copper nano-cutting, the strain
distribution is similar to that of the first group of bi-crystalline copper nano-cutting
when the tool is just cut into the workpiece, as shown in Fig. 20(b). However, when the
tool cuts through the grain boundary, the chip shape shows a pellet shape, which is
significantly different from the results of the first group, as shown in Fig. 20(b-1). In
the third bi-crystalline copper, when the tool is just cut into grain 2, only the contact
area between the tool and the workpiece generates concentrated strain, as shown in Fig.
20(c). Grain 2 induces grain 1 to produce a strain band dominated by crystal direction
of [-101] through grain boundaries. When the tool is cut across the grain boundary, the
large thickness of the banded chip is generated, and the obvious strain belt is left in
grain 2. One end of the strain belt is connected to the machined surface and the other
end is connected to the grain boundary, as shown in Fig. 20(c-1).
Fig. 20 Strain distribution of three groups of different bi-crystalline copper in cutting process. (a)
and (a-1): [100] [010] [001] with [210] [-120] [001]; (b) and (b-1): [100] [010] [001] with [110] [-
110] [001]; (c) and (c-1): [100] [010] [001] with [11-1] [112] [1-10].

Fig. 21 shows the cutting force variation diagram of three groups of double crystal
combinations in the nano-cutting process. It can be seen from the results that the value
of Fx and Fz are almost the same in the process of nano-cutting bi-crystalline copper
compared with polycrystalline copper. According to the strain cloud map of grain 1 in
Fig. 20, grain 1 was mainly subjected to strain along the direction of 45 ° during nano-
cutting process. Therefore, it can be considered that the tool is mainly subjected to the
resistance of the workpiece along the 45 ° direction during the nano-cutting process,
which may be the main reason that Fx and Fz similar in value. It can also be seen from
the results that the curve coincidence of Fx and Fz is the best in the cutting of second
group of bi-crystalline copper, no matter in the cutting of grain 1 or gain 2, as shown in
Fig. 21(b). In the first group of bi-crystalline copper nano-cutting, when grain 2 was
cut, the coincidence of Fx and Fz was poor, and the curve segment with the largest
difference was as high as 30 nN. When the tool began to cut into grain 1, the fitting
degree of Fx and Fz improved, as shown in Fig. 21(a). In the third group of bi-crystal
copper cutting, no matter in the cutting of grain 1 or grain 2, Fx and Fz can not obtain a
good coincidence, as shown in Fig. 21(c). This may be because the crystal orientation
difference of the third group of bi-crystal copper is too large, resulting in a phase
difference in the curve change between Fx and Fz.

Fig. 21 Cutting force variation of three groups of bi-crystal copper in cutting process. (a): [100]
[010] [001] with [210] [-120] [001]; (b): [100] [010] [001] with [110] [-110] [001]; (c): [100] [010]
[001] with [11-1] [112] [1-10].

3.7 Effect of texture on dislocation density


In fact, polycrystalline copper material commonly used in industry are often not
randomly oriented, but selectively oriented to form a certain texture. This is because
the grains preferentially arrange in a certain direction under external tensile or
compressive stress. Taking the plate texture as an example, the {112}<111> texture of
polycrystalline copper is easy to form after cold rolling due to the influence of
lamination fault energy. If it is recrystallized, polycrystalline copper is easy to form
{124}<211> or {236}<385> texture. The grain arrangement of different textures is
different, so different sizes of dislocations may be introduced in the nano-cutting
process, which have different degrees of influence on the surface integrity of the
machined surface. Therefore, it is of great significance to explore the influence of
different textures on the dislocation density in nano-cutting process to obtain
polycrystalline copper workpiece with good surface integrity. In this part, the
dislocation distribution and change of dislocation density of three different textures
{112}<111>, {124}<211> and {236}<385> texture in nano-cutting process were
studied. The possible types of Euler angles for each texture can be obtained by using
Eq. (1) and Eq. (2). Then, ATOMSK software can be used to generate the texture
polycrystalline model through the obtained Euler angles. Fig. 22 shows the (110) pole
diagram of these three plate textures and labels their Euler angles distribution. X is the
rolling direction of the plate texture, and Y is the direction perpendicular to the rolling
direction. In this section, the rake angle of the tool is 5 °, the cutting edge radius is 1
nm, the cutting depth is 1 nm, and the cutting speed is 100 m/s.
Fig. 22 (110) Pole diagram of three different polycrystalline copper textures. (a): Texture
{112}<111>; (b) Texture {124}<211>; (c) Texture {236}<385>.

Fig. 23 shows the chip shapes and dislocation distribution of three different textures
in nano-cutting. It can be seen from the results that the chip thickness of texture
{112}<111> is relatively large and presents a spheroid shape, as shown in Fig. 23(a).
The chip thickness of {124}<211> texture is smaller than that of {112}<111> texture
and the thickness difference between the chip tip and the chip middle area is larger,
which can be seen from the Fig. 23(b). The chip thickness of {236}<385> texture is the
smallest and shows the typical banded chip characteristics, as shown in Fig. 23(c). In
terms of dislocation, the dislocation line density of texture {112}<111> is higher than
that of texture {124}<211> and {236}<385>. Furthermore, the dislocation line density
of each grain varies greatly. For example, the dislocation lines of grain 1 is sparse, while
the dislocation line of grain 2 is dense, as shown in Fig. 23(a-1). In addition, the
dislocation line density between grain boundaries of texture {112}<111> is also high,
as shown by the dislocation line between grain 2 and grain 3. The dislocation line
density of texture {124}<211> is sparser when compared with texture {112}<111>.
Moreover, the dislocation line category within a single grain is single, as shown in grain
1 and 2 in Fig. 23(b-1). The bule dislocation line in the figure is perfect dislocation line,
the green dislocation line is Shockley dislocation line, and the red dislocation line is
unidentified dislocation line. The dislocation line density of texture {236}<385> is in
a moderate state compared with the previous two textures. In addition, the intertwining
state of dislocation lines in each grain of the texture is uniform, as shown in Fig. 23(c-
1).
In order to show the evolution law of dislocation density more clearly, Fig. 24 also
shows the change of the total dislocation density of the above three textures in the nano-
cutting process. The results show that texture {112}<111> has the largest dislocation
density, followed by texture {236}<385> and texture {124}<211> has the smallest
dislocation density. Moreover, it can also be seen that dislocation density will
periodically increase and decrease with increase of cutting distance, but the overall
trend is upward. Due to the small increase in the dislocation density of texture
{236}<385>, the dislocation density of texture {124}<211> exceeds that of texture
{236}<385> when the cutting distance reaches about 16 nm.

Fig. 23 Chip shape and dislocation distribution during nano-cutting of polycrystalline copper with
different textures. (a) and (a-1): Texture {112}<111>; (b) and (b-1): Texture {124}<211>; (c) and
(c-1): Texture {236}<385>.
0.0040
Texture
texture-1{112}<111>
texture-2{124}<211>
Texture
0.0035 texture-3{236}<385>
Texture

Dislocation density (nm/nm3)


0.0030

D
0.0025

0.0020

0.0015

4 6 8 10 12 14 16 18
Cutting distance (nm)

Fig. 24 Dislocation densities under different textures change with cutting distance.
4. Conclusions
In this study, the effects of cutting depth, tool geometry, grain size and grain
orientation on the material removal mechanism of polycrystalline copper nano-cutting
were investigated using molecular dynamics. The following conclusions can be draw
from the study of atomic migration, chip shape, dislocation density, strain distribution
and cutting forces in the nano-cutting process:
(1) With the increase of cutting depth, the number of HCP atoms representing stacking
dislocations in chip slip belt increases. The number of HCP atoms is positively
correlated with the cutting depth. In addition, when the cutting depth increases from 1
nm to 4 nm, the chip shear angle decreases from 45 ° to 35 °. The tool rake angle mainly
determines the chip thickness and curl degree, and the radius of cutting edge mainly
determines the atomic migration and strain distribution.
(2) Smaller grain size shows more accumulation of HCP-dominated stacking
dislocation. In addition, the smaller the grain size, the more concentrated the von Mises
stress distribution in the chip in front of the tool rake surface. Furthermore, the smaller
the grain size is, the more the hydrostatic pressure is distributed in multiple direction.
(3) The orientation difference affects chip shape, strain distribution and cutting force in
the bi-crystalline copper. The [11-1] [112] [1-10] grain only produces large plastic strain
in the contact area between the tool and workpiece, while the grains [210] [-120] [001]
and [110] [-110] [001] could produce large plastic strain in the whole grain.
(4) Different polycrystalline copper textures have different dislocation densities during
nano-cutting process. Texture {112}<111> has the highest dislocation density, followed
by texture {236}<385>, and texture {124}<211> has the lowest dislocation density.
Dislocation density fluctuates periodically with the increase of cutting distance during
the cutting process, but the overall trend is increase.

Declaration of Competing Interest


The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in this
paper.

Acknowledgements
The authors gratefully acknowledge the financial support from the National Natural
Science Foundation of China (Nos. 51775147 and 52005133).
References
1. Fan, P., et al., Rapid fabrication of surface micro/nano structures with enhanced broadband
absorption on Cu by picosecond laser. Optics Express, 2013. 21(10): p. 11628-11637.
2. Dai, Y., et al., Forced-based tool deviation induced form error identification in single-point
diamond turning of optical spherical surfaces. Precision Engineering, 2021. 72: p. 83-94.
3. Gawande, M.B., et al., Cu and Cu-based nanoparticles: synthesis and applications in catalysis.
Chemical reviews, 2016. 116(6): p. 3722-3811.
4. Walter, A.L., et al., Electronic structure of graphene on single-crystal copper substrates.
Physical Review B, 2011. 84(19): p. 195443.
5. Guo, J., et al., Kirkendall effect in creating three-dimensional metal catalysts for hierarchically
porous ultrathin graphite with unique properties. Chemistry of Materials, 2017. 29(11): p. 4991-
4998.
6. Lai, M., et al., Fundamental investigation on partially overlapped nano-cutting of
monocrystalline germanium. Precision Engineering, 2017. 49: p. 160-168.
7. Guo, S., et al., Microstructural evolution in ultra-precision grinding of Al/SiCp metal matrix
composites. Precision Engineering, 2023. 83: p. 12-21.
8. Shi, C., et al., Quasi-static kinematics model for motion errors of closed hydrostatic guideways
in ultra-precision machining. Precision Engineering, 2021. 71: p. 90-102.
9. Zhao, C., C.F. Cheung, and M. Liu, Nanoscale measurement with pattern recognition of an
ultra-precision diamond machined polar microstructure. Precision Engineering, 2019. 56: p.
156-163.
10. Hao, Z., Z. Lou, and Y. Fan, Study on staged work hardening mechanism of nickel-based single
crystal alloy during atomic and close-to-atomic scale cutting. Precision Engineering, 2021. 68:
p. 35-56.
11. Tao, Y., et al., High-accurate cutting forces estimation by machine learning with voice coil
motor-driven fast tool servo for micro/nano cutting. Precision Engineering, 2023. 79: p. 291-
299.
12. Wang, M.H., et al., Effect of dynamic adjustment of diamond tools on nano-cutting behavior of
single-crystal silicon. Applied Physics A, 2019. 125: p. 1-13.
13. Liu, B., et al., Experimental study on size effect of tool edge and subsurface damage of single
crystal silicon in nano-cutting. The International Journal of Advanced Manufacturing
Technology, 2018. 98: p. 1093-1101.
14. Dong, Z., et al., Effects of minimum uncut chip thickness on tungsten nano-cutting mechanism.
International Journal of Mechanical Sciences, 2023. 237: p. 107790.
15. Xu, D., et al., Revealing Nanoscale deformation mechanisms caused by shear-based material
removal on individual grains of a Ni-based superalloy. Acta Materialia, 2021. 212: p. 116929.
16. Kieren-Ehses, S., et al., The influence of the crystallographic orientation when micro machining
commercially pure titanium: a size effect. Precision Engineering, 2021. 72: p. 158-171.
17. Hu, G., W. Wu, and H. Dai, Atomistic study of nano-cutting bicrystal cubic silicon carbide.
Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical
Engineering Science, 2023. 237(7): p. 1706-1718.
18. Wang, H., et al., Effects of cutting force on formation of subsurface damage during nano-cutting
of single-crystal tungsten. Journal of Manufacturing Science and Engineering, 2022. 144(11):
p. 111008.
19. Chen, C., M. Lai, and F. Fang, Subsurface deformation mechanism in nano-cutting of gallium
arsenide using molecular dynamics simulation. Nanoscale Research Letters, 2021. 16: p. 1-10.
20. Sun, X. and K. Cheng, Multi-scale simulation of the nano-metric cutting process. The
International Journal of Advanced Manufacturing Technology, 2010. 47: p. 891-901.
21. Yang, S., et al., Multi-scale numerical analysis and experimental verification for nano-cutting.
Journal of Manufacturing Processes, 2021. 71: p. 260-268.
22. Liu, B., et al., Effect of ion implantation on material removal mechanism of 6H-SiC in nano-
cutting: A molecular dynamics study. Computational Materials Science, 2020. 174: p. 109476.
23. Liu, B., et al., In situ experimental study on material removal behaviour of single-crystal silicon
in nanocutting. International Journal of Mechanical Sciences, 2019. 152: p. 378-383.
24. Tian, D., et al., In situ investigation of nanometric cutting of 3C-SiC using scanning electron
microscope. The International Journal of Advanced Manufacturing Technology, 2021. 115(7-8):
p. 2299-2312.
25. Wang, Z., et al., Crystal anisotropy-dependent shear angle variation in orthogonal cutting of
single crystalline copper. Precision Engineering, 2020. 63: p. 41-48.
26. Shetty, N., et al., A review on finite element method for machining of composite materials.
Composite Structures, 2017. 176: p. 790-802.
27. Chan, C., W.B. Lee, and H. Wang, Enhancement of surface finish using water-miscible nano-
cutting fluid in ultra-precision turning. International Journal of Machine Tools and Manufacture,
2013. 73: p. 62-70.
28. Fang, F., et al., A study on mechanism of nano-cutting single crystal silicon. Journal of materials
processing technology, 2007. 184(1-3): p. 407-410.
29. Le, J.-B. and J. Cheng, Modeling electrified metal/water interfaces from ab initio molecular
dynamics: Structure and Helmholtz capacitance. Current Opinion in Electrochemistry, 2021.
27: p. 100693.
30. Cui, D.-D. and L.-C. Zhang, Nano-machining of materials: understanding the process through
molecular dynamics simulation. Advances in Manufacturing, 2017. 5: p. 20-34.
31. Brant, A. and M. Sundaram, Molecular dynamics study of direct localized overpotential
deposition for nanoscale electrochemical additive manufacturing process. Precision
Engineering, 2019. 56: p. 412-421.
32. Stefanou, G., The stochastic finite element method: past, present and future. Computer methods
in applied mechanics and engineering, 2009. 198(9-12): p. 1031-1051.
33. Guo, Z.X., Multiscale materials modelling: Fundamentals and applications. 2007: Elsevier.
34. Sun, X., et al., Multiscale simulation on nanometric cutting of single crystal copper. Proceedings
of the Institution of Mechanical Engineers, Part B: Journal of Engineering Manufacture, 2006.
220(7): p. 1217-1222.
35. Ye, Y., et al., Molecular dynamics simulation of nanoscale machining of copper.
Nanotechnology, 2003. 14(3): p. 390.
36. Tanaka, H., S. Shimada, and L. Anthony, Requirements for ductile-mode machining based on
deformation analysis of mono-crystalline silicon by molecular dynamics simulation. CIRP
annals, 2007. 56(1): p. 53-56.
37. Li, J., et al., A molecular dynamics investigation into plastic deformation mechanism of
nanocrystalline copper for different nanoscratching rates. Computational Materials Science,
2016. 118: p. 66-76.
View publication stats

38. Liu, H., et al., Molecular dynamics simulation of dislocation evolution and surface mechanical
properties on polycrystalline copper. Applied Physics A, 2019. 125: p. 1-13.
39. Liu, H., Y. Guo, and P. Zhao, Surface generation mechanism of monocrystalline materials under
arbitrary crystal orientations in nanoscale cutting. Materials Today Communications, 2020. 25:
p. 101505.
40. Tang, Yizhe, Eduardo M. Bringa, and Marc A. Meyers. Inverse Hall–Petch relationship in
nanocrystalline tantalum. Materials Science and Engineering: A 580 (2013): 414-426.
41. Naik, Sneha N., and Stephen M. Walley. The Hall–Petch and inverse Hall–Petch relations and
the hardness of nanocrystalline metals. Journal of Materials Science 55.7 (2020): 2661-2681.

You might also like