You are on page 1of 9

Material Strengthening

Mechanisms and Their


Contribution to Size Effect in
Micro-Cutting
Kai Liu
Shreyes N. Melkote
The George W. Woodruff School of Mechanical
Engineering,
Georgia Institute of Technology,
Atlanta, GA 30332-0406

The specific cutting energy in machining is known to increase nonlinearly with decrease
in uncut chip thickness. It has been reported in the literature that this phenomenon is
dependent on several factors such as material strengthening, ploughing due to finite edge
radius, and material separation effects. This paper examines the material strengthening
effect where the material strength increases nonlinearly as the uncut chip thickness is
reduced to a few microns. This increase in strength has been attributed in the past to
various factors such as strain rate, strain gradient, and temperature effects. Given that
the increase in material strength can occur due to many factors, it is important to
understand the contributions of each factor to the increase in specific cutting energy and
the conditions under which they are dominant. This paper analyzes two material strengthening factors, (i) the contribution of the decrease in the secondary deformation zone
cutting temperature and (ii) strain gradient strengthening, and their relative contributions to the increase in specific cutting energy as the uncut chip thickness is reduced.
Finite element (FE)-based orthogonal cutting simulations are performed with Aluminum
5083-H116, a work material with a small strain rate hardening exponent, thus minimizing
strain rate effects. Suitable cutting conditions are identified under which the temperature
and strain gradient effects are dominant. Orthogonal cutting experiments are used to
validate the model in terms of cutting forces. The simulation results are then analyzed to
identify the contributions of the material strengthening factors to the size effect in specific
cutting energy. DOI: 10.1115/1.2193548

Introduction
In the past decade, the need for miniaturized components in
fields such as MEMS, biomedical devices, and micro-molding has
increased. Advances in these fields have drawn the attention of
several researchers to micro-machining processes and material
properties at the micron and submicron levels. Mechanical microcutting is a promising manufacturing process that is capable of
producing micro-scale three-dimensional features with high accuracy and precision. However, the fundamental mechanics of mechanical micro-cutting is not well understood. For example, there
exists no consensus on a phenomenological explanation of scaling
effects in micro-cutting such as that observed in the specific cutting energy, where the amount of energy required for removing
unit volume of material increases nonlinearly when the uncut chip
thickness decreases from a few hundred microns to a few microns.
Many researchers have attempted to explain and predict this
scaling phenomenon, which is also termed the size effect. Shaw
1 and Backer et al. 2 attributed the size effect to a significantly
reduced number of imperfections encountered when deformation
takes place in a small volume. Therefore, material strength would
be expected to increase and approach the theoretical strength. It is
also well known that the flow stress of metals depends on the
strain, strain rate, and temperature. From dislocation mechanics,
material strength in plastic deformation of metallic crystals is determined by the motion of dislocations and their interactions. An
increase in temperature increases the thermodynamic probability
of the dislocations achieving sufficient energy to move past a peak
in the potential, thereby producing a softening effect on the flow
Contributed by the Manufacturing Engineering Division of ASME for publication
in the JOURNAL OF MANUFACTURING SCIENCE AND ENGINEERING. Manuscript received
June 9, 2005; final manuscript received November 30, 2005. Review conducted by
D.-W. Cho.

730 / Vol. 128, AUGUST 2006

stress. The rate of strain is proportional to the rate of dislocation


formation so that an increase in strain rate has a hardening effect
on the flow stress. Kopalinsky and Oxley 3 and Marusich 4
attributed the size effect in machining to an increase in the shear
strength of the workpiece material due to a decrease in the toolchip interface temperature at small uncut chip thickness values.
Larsen-Basse and Oxley 5 attributed the size effect to material
strengthening arising from an increase in the strain rate in the
primary shear zone with decrease in uncut chip thickness. Fang
6 recently presented a complex slip line model for orthogonal
machining and attributed the size effect to the material constitutive behavior of varying shear flow stress.
It was recently reported 79 that in nano-indentation the hardness of a metal is dependent on the indentation depth when it is of
the order of a few microns. Experimental evidence from microtwisting of thin copper wire 10 and micro-bending of thin nickel
beams 11 also display strong size effects when the characteristic
length associated with nonuniform plastic deformation is of the
order of a few microns. Dinesh et al. 12 suggested that the
material strengthening effect in micro-cutting processes is analogous to that in nano-indentation because of the intense localized
inhomogeneous plastic deformation within the primary and secondary deformation zones. Building on the work in 12, Joshi and
Melkote 13 presented an analytical model for orthogonal cutting
that incorporates a material constitutive law with strain gradient
dependence of flow stress. Liu and Melkote 14 have recently
developed a strain-gradient-based finite element model for micro/
meso-scale orthogonal cutting. Their simulation results predict a
sizeable strain gradient strengthening effect in micro-cutting.
Besides material strengthening, other reasons have been also
attributed to the size effect in specific cutting energy. Nakayama
and Tamura 15 conducted micro-cutting tests at a very low cutting speed 0.1 m / min to minimize temperature and strain rate

Copyright 2006 by ASME

Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

effects and still observed the size effect. They attributed the increase in specific cutting energy to a decrease in the shear plane
angle and greater energy dissipation associated with plastic flow
in the workpiece subsurface. Kim et al. 16 and Lucca et al. 17
considered the tool edge radius as the major cause of size effect.
Armarego and Brown 18 suggested that the increase in specific
cutting force with decrease in undeformed chip thickness was due
to greater relative contribution of ploughing forces arising from
frictional rubbing and ploughing associated with material removal
by a blunt tool. Komanduri 19 reported size effect due to the tool
edge radius at nanometer length scales by carrying out molecular
dynamics simulations of orthogonal cutting at depths of cut of
0.362 12.172 nm and tool edge radii of 3.62 21.72 nm. Atkins
20 attributed the size effect in cutting to the energy required for
new surface creation via ductile fracture. This energy is thought to
be independent of the depth of cut and, consequently, its contribution to the overall work increases at small depths of cut.
It is clear from the literature that material strengthening at small
uncut chip thickness values is an important contributor to the size
effect in specific cutting energy. While material strengthening has
been attributed to several factors, it is not clear what the contribution of each factor is and under what conditions their effect is
dominant. Therefore, to partially address this need, this paper focuses on two main strengthening factors: i the contribution of
the decrease in the secondary deformation zone cutting temperature with decrease in uncut chip thickness and ii strain gradient
strengthening. Finite element FE simulations of orthogonal
micro-cutting are performed using aluminum 5083-H116, a material with a small strain rate hardening exponent, thus minimizing
strain rate effects. Suitable cutting conditions are identified under
which the temperature and strain gradient effects are dominant.
Orthogonal cutting experiments are used to validate the model
using cutting forces. The simulation results are then analyzed to
identify the contributions of these two material strengthening factors to the size effect in specific cutting energy.

The density of statistically stored dislocation, s, can be determined from the uniaxial stress-strain law in the absence of strain
gradient effects as,

= 3Gbs = ref f

where ref is the reference stress in uniaxial tension. On the other


hand, the density of geometrically necessary dislocations, g, can
be related to the effective strain gradient as,

g = 2/b

Based on these equations, a flow stress equation accounting for


the effect of geometrically necessary dislocations can be written
as,

= ref f 2, ,T* + l

where l is the material length scale and is given as,


l = 182


G
ref

The flow stress function, f, is assumed to be of the JohnsonCook 22 form as follows:

f, ,T* = A + Bn 1 + C ln

1 T*m
o

The stain gradient tensor is defined as,

ijk = uk,ij = ik,j + jk,i ij,k

The effective strain gradient, which measures the density of


geometrically necessary dislocations, is defined as,

ijk

= 41 ijk

where the third-order deviatoric strain gradient tensor ijk is given


by,

= ijk 41 ik jpp + jkipp


ijk

Model Formulation
In this paper, a strain-gradient-based finite element model for
micro/meso-scale orthogonal cutting processes developed earlier
by Liu and Melkote 14 is adopted as the simulation platform.
The FE model is a fully coupled thermal-mechanical model developed in the commercially available software, ABAQUS/
Standard version 6.4. This section reviews the following aspects
of the finite element model: a constitutive model, b modeling
of tool-chip interaction, c modeling of chip separation, and d
modeling of the heat transfer. Several key techniques and the
overall simulation approach are reviewed briefly.
Constitutive Modeling. A Taylor-based nonlocal theory of
plasticity proposed by Gao and Huang 21 is used to represent
the material behavior under highly localized inhomogeneous deformation. From the viewpoint of dislocation mechanics, the material is work hardened due to the formation, motion, and interaction of dislocations. Statistically stored dislocations accumulate by
trapping each other in a random way while geometrically necessary dislocations are required for compatible deformation and are
related to the gradient of plastic strain. This theory links Taylors
model of dislocation hardening to a nonlocal theory of plasticity
in which the density of geometrically necessary dislocations is
expressed as a nonlocal integral of the strain field. Preserving the
structure of classical continuum theory, the balance law of the
theory, i.e., the balance of angular and linear momentum, is identical to the classical theories. The Taylor dislocation model defines
the shear flow stress in terms of the dislocation density as
= Gb, where the dislocation density is composed of the density of statistically stored dislocations, s, and the geometrically
necessary dislocations, g.
Journal of Manufacturing Science and Engineering

Using Taylor series expansion of the strain components 21,


the deviatoric strain gradient ijk is obtained from the following
nonlocal volume integral,

=
ijk

ik j + jki ijk 4 ik j
1

+ jki ppd


m k d

In the current finite element implementation of this model, the


above integral is reduced to an area integral due to orthogonal
cutting assumption and is evaluated over a mesoscale cell defined
around each Gaussian integration point and contained within the
corresponding element 23.
The constitutive equations of Taylor-based nonlocal theory of
plasticity 21 are expressed in rate form as follows:

kk = 3k kk
ij = 2 ij

if e = and e 0

= 2G ij

2
l
3ij 6ij ij ref
2
4e 23e + ref f f

10

if e or e 0
11

The key feature of the Taylor-based nonlocal theory of plasticity is that it does not involve higher order terms and preserves the
structure of classical continuum mechanics. Strain gradient enters
the constitutive model as a nonlocal integral and affects the flow
stress variation. Thus, it has the advantage of simpler implementation compared to other gradient plasticity theories 2426.
AUGUST 2006, Vol. 128 / 731

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

Choice of Finite Element Platform

Fig. 1 Illustration of material separation using the pure deformation method

Tool-Chip Interaction. Accurate representation of the interaction between the tool and chip is vital for obtaining a reliable and
realistic simulation. However, the friction characteristic at the tool
chip interface is difficult to determine since it is influenced by
many factors such as cutting speed, contact pressure, and temperature. Extensive studies have been performed on the mechanics of
interaction along the tool-chip interface and several models have
been developed. Of these, Zorevs model 27, which reveals that
two distinct regions of sliding and sticking on the interface exist,
is widely accepted. In the sliding region, the shear stress is a
fraction of the normal contact pressure, p. As the shear stress
reaches a limiting shear stress value, *, sticking occurs and the
shear stress equals the limiting shear stress value regardless of the
normal contact pressure. The extended Coulomb friction model,
expressed in terms of the frictional shear stress see Eq. 12,
appears to fit the machining problem adequately and has been
used successfully by several researchers 2833 and is chosen in
this work to model the tool-chip interaction.
s = p,

when p * sliding

s = *,

when p * sticking

12

Chip Separation Modeling. The pure deformation method of


chip formation 4,14,34,35 has been implemented in this work
with the help of the adaptive remeshing technique, thus avoiding a
chip separation criterion. The cutting process is likened to a metal
forming operation, with material flowing on the two sides of the
tool. There is no predefined parting line and, therefore, the shape
of the chip is not predetermined. Instead, as the tool advances,
nodes in the workpiece move along the tool surface, causing the
elements to deform severely near the tool tip. The severely distorted elements are replaced during the remeshing step by new
elements that are more regular in shape. The material that overlaps
the tool is removed during the remeshing step see Fig. 1.
Heat Transfer Model. In the finite element model, heat generation due to plastic deformation and friction at the tool-chip
interface is modeled as a volume heat flux. Heat conduction is
assumed to be the primary mode of heat transfer, which occurs
within the workpiece material and at the tool-chip interface.
The governing equation of heat transfer is as follows:
K

T
2T
2T
T

+
K
= mC p u +
+Q
x2
y2
x
y

13

where K is the thermal conductivity of the workpiece material, m


is the mass density, C p is the specific heat capacity, u is the ve
locity in the x direction, v is the velocity in the y direction, and Q
is the volume heat flux.
The fraction of dissipated energy converted into heat due to
plastic deformation and friction is assumed to be 0.9. Heat generated due to friction is distributed via a weighting factor of 0.5
between the two contact surfaces. Thermal boundary conditions
elsewhere are set to zero.
732 / Vol. 128, AUGUST 2006

There are many commercial finite element codes that are


capable of modeling the machining process. Examples include
DEFORM 2D/3D, ABAQUS, and Third Waves AdvantEdge Classic. Based on the modeling requirements noted above,
the finite element package required for the present work must
have the important feature of adaptive remeshing. A few of the
aforementioned packages such as DEFORM, AdvantEdge,
and MARC 2D possess this feature. However, the program
should also provide the user with the flexibility to define a genuine strain gradient plasticity material model using higher order
elements. Higher order elements are necessary for fully evaluating
the strain gradient tensor since the constant strain components in
first-order elements yield zero strain gradient. Although
ABAQUS/Standard does not have true adaptive remeshing capability, it is more flexible than some of the other codes as it
allows user-defined subroutines and higher order elements. Consequently, the general FEA package ABAQUS/Standard was selected as the finite element platform for this work.

Choice of Element Type


A coupled temperature-displacement plane strain element is required for simulating the orthogonal metal cutting process. For a
fully coupled simulation, ABAQUS only allows fully integrated
elements. However, as explained above, higher order elements are
required for the evaluation of strain gradient at the element integration points. Therefore, an eight-node biquadratic displacement
and bilinear temperature element was chosen to approximate the
workpiece geometry.

Adaptive Remeshing
Adaptive remeshing was implemented to avoid convergence
difficulties typically caused by severely distorted elements. The
remeshing module consisted of a preprocessor coded in
FORTRAN 77 plus the automatic mesh generator feature of ANSYS.
Interference depth is used as the remeshing criterion in the
model. During simulation, the amount of penetration between the
tool and workpiece contact pair is checked at each time step to
determine the interference depth. Once the remeshing criterion is
satisfied, the outline of the workpiece is stored and the automatic
mesh generating module of ANSYS is used to create a new
mesh for this region. Subsequently, the solution is mapped from
the old mesh to the new mesh.
It is necessary to have a very fine mesh in the primary and
secondary deformation zones to resolve the relatively steep stress
and strain gradients present in these zones. An element size of
2 m was used in these zones. However, using the same mesh
size throughout the workpiece increases the computational cost
significantly. Therefore, the mesh pattern generated in each remeshing step was designed to be much denser in the vicinity of
the two deformation zones, and coarser away from these zones.
A mesh density windowing technique was used for mesh refinement. As seen in Fig. 2, different mesh density windows were
defined in and away from the major deformation zones. This approach reduces the number of elements by a factor of 10 or more
and greatly reduces the computational cost while maintaining a
sufficiently high resolution for the solution.
The typical workpiece dimensions used in the simulations are
3 mm by 1 mm. As seen in Fig. 2, the bottom of the workpiece is
fully constrained while the top and right sides of the tool are fixed
in the Y direction. A velocity load is applied to the tool to simulate
the cutting speed.

Solution Mapping Scheme


The solution and state variables need to be mapped from the old
mesh to the newly created mesh after each remeshing step. Discontinuity in the solution is inevitable because of changes in the
Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

Acknowledgment
This work was supported by the National Science Foundation
through Grant No. DMI-0300457.

Nomenclature

o
p
l

s
*
p
i

Q
T*

T0
Tm
A
B
c
n
m
m
Cp
K
wx , y

ref

Fig. 16 Variation of maximum temperature in the primary and


secondary shear zones at 240 m / min cutting speed

seen that the specific cutting energy predicted by the model with
strain gradient almost doubles when the uncut chip thickness decreases from 10 to 0.5 m, while only about 10% increase is captured by the model without strain gradient effect. The plot shows
that at a high cutting speed and small uncut chip thickness, strain
gradient strengthening is more significant than material strengthening due to a drop in the secondary shear zone temperature.
It is evident from Fig. 16 that the temperature drop at the toolchip interface is less prominent at high cutting speeds and small
uncut chip thickness, which suggests that the temperature effect
contributes to only a small fraction of the size effect in microcutting under such conditions.

Conclusions
With the aim of understanding the contributions of different
material strengthening factors to the size effect in micro-cutting,
this paper focused primarily on two main strengthening factors: i
strain gradient strengthening and ii decrease in the secondary
deformation zone cutting temperature, with decrease in uncut chip
thickness. These factors were analyzed using a strain-gradientplasticity-based finite element model of orthogonal micro cutting
that was verified experimentally for aluminum alloy 5083-H116, a
material with a small strain rate hardening exponent, thus minimizing strain-rate effects. The model was then used to simulate
the size effect in microcutting under conditions where the temperature and strain gradient effects are dominant individually. The
following conclusions can be drawn from this work:

The strain-gradient-plasticity-based model of micro-cutting


is able to capture the size effect in specific cutting energy for
the aluminum alloy examined in this paper.
Strain gradient strengthening contributes significantly to the
size effect at low cutting speed and small uncut chip thickness 10 m.
Temperature dependence of flow stress plays a dominant
role in causing size effect at relatively high cutting speeds
and large uncut chip thickness. The size effect is caused by
material strengthening due to a drop in the secondary shear
zone temperature.
Strain gradient strengthening is more dominant than the
temperature effect at high cutting speed and small uncut
chip thickness. This implies that it is necessary to consider
the strain gradient effect, especially at the micron/submicron
levels.

Although the above conclusions were obtained from the study


of a strain rate insensitive material, similar results are expected for
other materials that are strain rate sensitive, since the strain gradient is determined by the spatial distribution of strain and not by
the strain rate. The existence and importance of strain gradient
shown in this paper suggests that it is necessary to consider this
effect when modeling micro-cutting processes.
Journal of Manufacturing Science and Engineering

ij

s
g
G

ijk
ijk

px , y


X

Xi

Burgers vector nm
flow stress MPa
effective stress
deviatoric stress
reference yield stress MPa
Kronecker tensor
dislocation density
statistically stored dislocation density
density of geometrically necessary dislocations
shear modulus
empirical constant
effective strain gradient
strain gradient tensor
deviatoric strain gradient tensor
effective true strain
effective strain rate
reference strain rate
effective plastic strain
characteristic material length
friction coefficient
frictional shear stress
limiting shear stress in Coulomb friction model
contact pressure
local coordinates in mesoscale cell
volume of the mesoscale cell
volume heat flux
dimensionless temperature in Johnson-Cook
equation= T T0 / Tm T0
ambient temperature
melting temperature
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material density
specific heat capacity
thermal conductivity
weighting function in diffuse approximation
method
polynomial basis in diffuse approximation
method
coefficient controlling the shape of weighting
function
position vector at point x
position vector at point xi

References
1 Shaw, M. C., 1950, A Quantized Theory of Strain Hardening as Applied to
the Cutting of Metals, J. Appl. Phys., 21, pp. 599606.
2 Backer, W. R., Marshall, E. R., and Shaw, M. C., 1952, The Size Effect in
Metal Cutting, Trans. ASME, pp. 7461.
3 Kopalinsky, E. M., and Oxley, P. L. B., 1984, Size Effects in Metal Removal
Processes, Proceedings of the Third Conference on Mechanical Properties at
High Rates of Strain Institute of Physics Conference Series, No. 70, Oxford,
UK, pp. 389396.
4 Marusich, T. D., 2001, Effects of Friction and Cutting Speed on Cutting
Force, Proc. IMECE ASME, Nov. 1116, New York, Paper No. MED23313.
5 Larsen-Basse, J., and Oxley, P. L. B., 1973, Effect of Strain Rate Sensitivity
on Scale Phenomena in Chip Formation, Proceedings of the 13th International Machine Tool Design & Research Conference, University of Birmingham, pp. 209216.

AUGUST 2006, Vol. 128 / 737

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

6 Fang, N., 2003, Slip-Line Modeling of Machining With a Rounded-Edge


ToolPart II: Analysis of the Size Effect and Shear Strain-Rate, J. Mech.
Phys. Solids, 51, pp. 743762.
7 Nix, W. D., 1989, Mechanical Properties of Thin Films, Metall. Trans. A,
20A, pp. 22172245.
8 Ma, Q., and Clarke, D. R., 1995, Size Dependent Hardness of Silver Single
Crystals, J. Mater. Res., 10, pp. 853863.
9 Stelmashenko, N. A., Walls, M. G., Brown, L. M., and Milman, Y. V., 1993,
Microindentation on W and Mo Oriented Single Crystals: an STM Study,
Acta Metall. Mater., 41, pp. 28552865.
10 Fleck, N. A., Muller, G. M., Ashby, M. F., and Hutchinson, J. W., 1994,
Strain Gradient Plasticity: Theory and Experiments, Acta Metall. Mater.,
422, pp. 475487.
11 Stolken, J. S., and Evans, A. G., 1998, A Microbend Test Method for Measuring the Plasticity Length Scale, Acta Mater., 4614, pp. 51095115.
12 Dinesh, D., Swaminathan, S., Chandrasekar, S., and Farris, T. N., 2001, An
Intrinsic Size-Effect in Machining Due to the Strain Gradient, Proc. 2001
ASME IMECE, Nov. 1116, New York, pp. 197204.
13 Joshi, S. S., and Melkote, S. N., 2004, An Explanation for the Size-Effect in
Machining Based on Strain Gradient Plasticity, ASME J. Manuf. Sci. Eng.,
1264, pp. 679684.
14 Liu, K., and Melkote, S. N., 2004, A Strain Gradient Based Finite Element
Model for Micro/Meso-Scale Orthogonal Cutting Process, Proceedings of
2004 Japan-USA Symposium on Flexible Automation, Denver, CO, July.
15 Nakayama, K., and Tamura, K., 1968, Size Effect in Metal-Cutting Force,
American Society of Mechanical Engineers, 67-Prod-9, 1967.
16 Kim, K. W., Lee, W. Y., and Sin, H. C., 1999, A Finite Element Analysis of
Machining With the Tool Edge Considered, J. Mater. Process. Technol., 86,
pp. 4555.
17 Lucca, D. A., Rhorer, R. L., and Komanduri, R., 1993, Effect of Tool Edge
Geometry on Energy Dissipation in Ultraprecision Machining, CIRP Ann.,
421, pp. 8386.
18 Armarego, E. J. A., and Brown, R. H., 1962, On the Size Effect in Metal
Cutting, Int. J. Prod. Res., 13, pp. 7599.
19 Komanduri, R., Chandrasekaran, N., and Raff, L. M., 1998, Effect of Tool
Geometry in Nanometric Cutting: a Molecular Dynamics Simulation Approach, Wear, 219, pp. 8497.
20 Atkins, A. G., 2003, Modelling Metal Cutting Using Modern Ductile Fracture
Mechanics: Quantitative Explanations for Some Longstanding Problems, Int.
J. Mech. Sci., 45, pp. 373396.
21 Gao, H., and Huang, Y., 2001, Taylor-Based Nonlocal Theory of Plasticity,
Int. J. Solids Struct., 38, pp. 26152637.
22 Johnson, G. R., and Cook, W. H., 1983, A Constitutive Model and Data for
Metals Subjected to Large Strain, High Strain Rates and High Temperatures,
Proc. 7th Int. Symposium on Ballistics, April, The Hague, Netherlands, pp.
541547.
23 Liu, K., 2005, Process Modeling of Micro-Cutting Including Strain Gradient
Effect, Ph.D thesis, Mechanical Engineering, Georgia Institute of Technology.
24 Fleck, N. A., and Hutchinson, J. W., 1993, A Phenomenological Theory for
Strain Gradient Effects in Plasticity, J. Mech. Phys. Solids, 41, pp. 1825
1857.
25 Gao, H., Huang, Y., Nix, W. D., and Hutchinson, J. W., 1999, Mechanism-

738 / Vol. 128, AUGUST 2006

26
27
28

29
30
31

32
33
34
35
36
37
38

39
40
41
42

Based Strain Gradient PlasticityI. Theory, J. Mech. Phys. Solids, 47, pp.
12391263.
Acharya, A., and Bassani, J. L., 2000, Lattice Incompatibility and a Gradient
Theory of Crystal Plasticity, J. Mech. Phys. Solids, 48, pp. 15651595.
Zorev, N. N., 1963, Inter-Relationship Between Shear Processes Occurring
Along Tool Face and Shear Plane in Metal Cutting, International Research in
Production Engineering, pp. 4249.
Usui, E., and Shirakashi, T., 1982, Mechanics of MachiningFrom Descriptive to Predictive Theory. In on the Art of Cutting Metals75 Years Later,
Proceedings of International Conference on Production Engineering Research,
September 912, Pittsburgh, PA, ASME, New York, PED-7, pp. 1335.
Shih, A. J., Chandrasekar, S., and Yang, H. T., 1990, Finite Element Simulation of Metal Cutting Process With Strain-Rate and Temperatures Effects,
Fundamental Issues in Machining, ASME, PED, 43, pp. 1124.
Bhattacharya, S., and Lovell, M. R., 2000, Characterization of Friction in
Machining: Evaluation of Asperity Deformation and Seizure-Based Models,
Trans. NAMRC/SME, XXVII, pp. 107112.
Lovell, M. R., Bhattacharya, S., and Zeng, R., 1998, Modeling Orthogonal
Machining for Variable Tool-Chip Interfacial Friction Using Explicit Dynamic
Finite Element Methods, Proceedings of the CIRP International Workshop on
Modeling of Machining Operations, Atlanta, GA, pp. 265276.
Tao, Z., and Lovell, M. R., 2002, Towards an Improved Friction Model in
Material Removal Processes: Investigating the Role of Temperature, SME
Technical Paper No. MR02-188, pp. 18.
Shi, G., Deng, X., and Shet, C., 2002, A Finite Element Study of the Effect of
Friction in Orthogonal Metal Cutting, Finite Elem. Anal. Design, 38, pp.
863883.
Madhavan, V., Chandrasekar, S., and Farris, T. N., 2000, Machining as a
Wedge Indentation, ASME J. Appl. Mech., 67, pp. 128139.
Ceretti, E., Fallbohmer, P., Wu, W. T., and Altan, T., 1996, Application of 2D
FEM to Chip Formation in Orthogonal Cutting, J. Mater. Process. Technol.,
59, pp. 169180.
Hamel, V., Roelandt, J. M., Gacel, J. N., and Schmidt, F., 2000, Finite Element Modeling of Clinch Forming With Automatic Remeshing, Comput.
Struct., 77, pp. 185200.
Fang, N., 2003, Sensitivity Analysis of the Material Flow Stress in Machining, American Society of Mechanical Engineers, Manufacturing Engineering
Division, MED, 14, pp. 2332.
Clausen, A. H., Borvik, T., Hopperstad, O. S., and Benallal, A., 2004, Flow
and Fracture Characteristic of Aluminum Alloy AA5083-H116 as Function of
Strain Rate, Temperature and Triaxiality, Mater. Sci. Eng., A, 364, pp. 260
272.
Altan, T., and Bougler, F. W., 1973, Flow Stress of Metals and its Application
in Metal Forming Analyses, American Society of Mechanical Engineers, No.
73, Prod-4.
Oxley, P. L. B., 1989, Mechanics of Machining: An Analytical Approach to
Assessing Machinability, Ellis Horwood Ltd., Chichester.
Zhou, M., and Clode, M. P., 1998, Constitutive Equations for Modeling Flow
Softening Due to Dynamic Recovery and Heat Generation During Plastic Deformation, Mech. Mater., 27, pp. 6376.
MatWeb material database http://www.matweb.com.

Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

Fig. 9 Comparison of experimental and simulated cutting


forces at 200 m / min

Fig. 8 Flow stress data for Al5083-H116 41

f, ,T* = A + Bn1 + c1 T*m

16

As seen from Figs. 9 and 10, good agreement is obtained between the predicted and measured forces at both cutting speeds.
At 200 m / min, the absolute average percentage error in the cutting force prediction is 4.25% and 23.21% for the thrust force
prediction. At 10 m / min, the absolute average percentage error in
the cutting force prediction is 9.6% and 22.2% for the thrust force
prediction. It can be seen that a better match is obtained for the
cutting force while the thrust force is generally underpredicted at
small uncut chip thicknesses for both cutting speeds. This could
be caused by the use of a constant friction coefficient for the
whole range of uncut chip thickness whereas the actual friction
coefficient estimated from the measured forces shows an increasing trend with decrease in uncut chip thickness. Nevertheless, the model gives fairly good predictions in terms of the absolute error. Thus, the finite element model is considered to be

satisfactorily validated and is used in the following section to


analyze the contribution of each material strengthening factor to
the size effect in micro-cutting.

Results And Discussion


Figure 11 shows the effective strain gradient distribution obtained from the simulation at 1 m depth of cut. It is found that
very high strain gradients are present at the tool-chip interface and

Table 3 Modified Johnson-Cook flow stress model coefficients for Al5083-H116


A MPa

B MPa

167

300

0.12

0.859

Table 4 Material properties of Al5083-H116 42


Density Kg/ m3
Specific heat J/kg C
Thermal conductivity
W/m K
Coefficient of thermal
expansion m / m C
Melting temperature C
Yield strength MPa
Youngs modulus GPa
Shear modulus GPa
Poissons ratio
Burgers vector nm

Fig. 10 Comparison of experimental and simulated cutting


forces at 10 m / min

2660
900
117
12.6
591638
228
71
26.4
0.33
0.256

Table 5 Material properties of diamond tools 42


Density Kg/ m3
Specific heat J/kg C
Thermal conductivity
W/m k
Coefficient of thermal
expansion m / m C
Melting temperature C
Youngs modulus GPa
Poissons ratio

3500
471.5
1500
2.0
4027
850
0.1

Journal of Manufacturing Science and Engineering

Fig. 11 Strain gradient contour uncut chip thickness 1 m,


10 m / min cutting speed

AUGUST 2006, Vol. 128 / 735

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

Fig. 14 Variation of maximum temperature in the primary and


secondary shear zones at 200 m / min cutting speed PSZ: primary shear zone, SSZ: secondary shear zone
Fig. 12 Variation of specific cutting energy with uncut chip
thickness at 10 m / min

in the workpiece surface layers. As expected, a high strain gradient band is also present in the primary deformation zone. Analysis
of the simulation results shows that the normal plastic strain components 11 and 22 not shown contribute to the strain gradient in
the primary deformation zone and the tool-chip interface, while
the high strain gradient in the surface layers comes mainly from
the distribution of shear plastic strain 12 not shown. With an
effective strain gradient of several hundred and a calculated material length scale of 5.7 m obtained by substituting the material
parameters into Eq. 4, it can be seen that the term associated
with the strain gradient in Eq. 3 is significant compared to the
other term. Consequently, the strain gradient strengthening effect
is considerable. This also suggests that the effect will become
even more dominant at smaller depths of cut, which will produce
an even steeper strain gradient.
In order to examine the strain gradient effect at a low cutting
speed of 10 m / min, two sets of orthogonal cutting simulations
were run at uncut chip thickness values ranging from
0.5 to 10 m. The first set of simulations was run with all terms
in Eq. 3 while the second set of simulations was run without the
strain gradient term l in Eq. 3. Figure 12 shows a plot of the
specific cutting energy versus uncut chip thickness with and without the strain gradient effect. The specific cutting energy was
computed by dividing the total force acting on the tool in the
cutting direction by the product of the workpiece width unity
and uncut chip thickness. It can be seen that the specific cutting
energy predicted by the model with strain gradient effect matches
well with the experimental data and captures the size effect. For
the model without strain gradient effect, the predicted specific
cutting energy remains constant with decrease in uncut chip thickness. It is clear from Fig. 12 that at a low cutting speed and for
small uncut chip thickness, strain gradient strengthening is the
dominant mechanism responsible for the size effect.

Fig. 13 Variation of specific cutting energy with uncut chip


thickness at 200 m / min

736 / Vol. 128, AUGUST 2006

Similarly, in order to examine the temperature effect at a cutting


speed of 200 m / min, two sets of orthogonal cutting simulations
were run from 20 to 200 m uncut chip thickness. The first set of
simulations was run with all terms in Eq. 3 while the second set
of simulations was run without the temperature term in Eq. 16
and hence in Eq. 3. Figure 13 shows a plot of the specific
cutting energy versus uncut chip thickness with and without the
temperature effect. It can be seen that the specific cutting energy
predicted by the model with the temperature effect is in good
agreement with the experimental data and captures the size effect.
Simulations without the temperature effect show that the specific
cutting energy remains fairly constant with reduction in uncut chip
thickness. Clearly, at the higher cutting speed the temperature effect is dominant compared to the strain gradient effect, especially
at large uncut chip thickness values.
The maximum temperatures in the primary and secondary shear
zones versus uncut chip thickness are shown in Fig. 14. It can be
seen that the maximum temperature in the secondary shear zone
drops by nearly 200 C while the maximum temperature in the
primary shear zone remains almost unchanged with a decrease in
uncut chip thickness from 200 to 20 m. This directly supports
the reasoning proposed by Kopalinsky and Oxley 3 and Marusich 4 that the size effect in machining at high cutting speeds
and large uncut chip thickness is primarily caused by an increase
in the shear strength of the workpiece material due to a decrease
in the tool-chip interface temperature.
It is also of interest to examine the contributions of the two
material strengthening factors at cutting conditions characterized
by high cutting speed and small uncut chip thickness. Under such
conditions, both temperature and strain gradient are expected to
contribute to the size effect in micro-cutting. Therefore, orthogonal cutting simulations were run at a cutting speed of 240 m / min
for uncut chip thickness values ranging from 0.5 to 10 m. Figure 15 shows a plot of the specific cutting energy versus uncut
chip thickness with and without strain gradient effect. It can be

Fig. 15 Variation of specific cutting energy with uncut chip


thickness at 240 m / min

Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

Acknowledgment
This work was supported by the National Science Foundation
through Grant No. DMI-0300457.

Nomenclature

o
p
l

s
*
p
i

Q
T*

T0
Tm
A
B
c
n
m
m
Cp
K
wx , y

ref

Fig. 16 Variation of maximum temperature in the primary and


secondary shear zones at 240 m / min cutting speed

seen that the specific cutting energy predicted by the model with
strain gradient almost doubles when the uncut chip thickness decreases from 10 to 0.5 m, while only about 10% increase is captured by the model without strain gradient effect. The plot shows
that at a high cutting speed and small uncut chip thickness, strain
gradient strengthening is more significant than material strengthening due to a drop in the secondary shear zone temperature.
It is evident from Fig. 16 that the temperature drop at the toolchip interface is less prominent at high cutting speeds and small
uncut chip thickness, which suggests that the temperature effect
contributes to only a small fraction of the size effect in microcutting under such conditions.

Conclusions
With the aim of understanding the contributions of different
material strengthening factors to the size effect in micro-cutting,
this paper focused primarily on two main strengthening factors: i
strain gradient strengthening and ii decrease in the secondary
deformation zone cutting temperature, with decrease in uncut chip
thickness. These factors were analyzed using a strain-gradientplasticity-based finite element model of orthogonal micro cutting
that was verified experimentally for aluminum alloy 5083-H116, a
material with a small strain rate hardening exponent, thus minimizing strain-rate effects. The model was then used to simulate
the size effect in microcutting under conditions where the temperature and strain gradient effects are dominant individually. The
following conclusions can be drawn from this work:

The strain-gradient-plasticity-based model of micro-cutting


is able to capture the size effect in specific cutting energy for
the aluminum alloy examined in this paper.
Strain gradient strengthening contributes significantly to the
size effect at low cutting speed and small uncut chip thickness 10 m.
Temperature dependence of flow stress plays a dominant
role in causing size effect at relatively high cutting speeds
and large uncut chip thickness. The size effect is caused by
material strengthening due to a drop in the secondary shear
zone temperature.
Strain gradient strengthening is more dominant than the
temperature effect at high cutting speed and small uncut
chip thickness. This implies that it is necessary to consider
the strain gradient effect, especially at the micron/submicron
levels.

Although the above conclusions were obtained from the study


of a strain rate insensitive material, similar results are expected for
other materials that are strain rate sensitive, since the strain gradient is determined by the spatial distribution of strain and not by
the strain rate. The existence and importance of strain gradient
shown in this paper suggests that it is necessary to consider this
effect when modeling micro-cutting processes.
Journal of Manufacturing Science and Engineering

ij

s
g
G

ijk
ijk

px , y


X

Xi

Burgers vector nm
flow stress MPa
effective stress
deviatoric stress
reference yield stress MPa
Kronecker tensor
dislocation density
statistically stored dislocation density
density of geometrically necessary dislocations
shear modulus
empirical constant
effective strain gradient
strain gradient tensor
deviatoric strain gradient tensor
effective true strain
effective strain rate
reference strain rate
effective plastic strain
characteristic material length
friction coefficient
frictional shear stress
limiting shear stress in Coulomb friction model
contact pressure
local coordinates in mesoscale cell
volume of the mesoscale cell
volume heat flux
dimensionless temperature in Johnson-Cook
equation= T T0 / Tm T0
ambient temperature
melting temperature
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material constant in Johnson-Cook equation
material density
specific heat capacity
thermal conductivity
weighting function in diffuse approximation
method
polynomial basis in diffuse approximation
method
coefficient controlling the shape of weighting
function
position vector at point x
position vector at point xi

References
1 Shaw, M. C., 1950, A Quantized Theory of Strain Hardening as Applied to
the Cutting of Metals, J. Appl. Phys., 21, pp. 599606.
2 Backer, W. R., Marshall, E. R., and Shaw, M. C., 1952, The Size Effect in
Metal Cutting, Trans. ASME, pp. 7461.
3 Kopalinsky, E. M., and Oxley, P. L. B., 1984, Size Effects in Metal Removal
Processes, Proceedings of the Third Conference on Mechanical Properties at
High Rates of Strain Institute of Physics Conference Series, No. 70, Oxford,
UK, pp. 389396.
4 Marusich, T. D., 2001, Effects of Friction and Cutting Speed on Cutting
Force, Proc. IMECE ASME, Nov. 1116, New York, Paper No. MED23313.
5 Larsen-Basse, J., and Oxley, P. L. B., 1973, Effect of Strain Rate Sensitivity
on Scale Phenomena in Chip Formation, Proceedings of the 13th International Machine Tool Design & Research Conference, University of Birmingham, pp. 209216.

AUGUST 2006, Vol. 128 / 737

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

6 Fang, N., 2003, Slip-Line Modeling of Machining With a Rounded-Edge


ToolPart II: Analysis of the Size Effect and Shear Strain-Rate, J. Mech.
Phys. Solids, 51, pp. 743762.
7 Nix, W. D., 1989, Mechanical Properties of Thin Films, Metall. Trans. A,
20A, pp. 22172245.
8 Ma, Q., and Clarke, D. R., 1995, Size Dependent Hardness of Silver Single
Crystals, J. Mater. Res., 10, pp. 853863.
9 Stelmashenko, N. A., Walls, M. G., Brown, L. M., and Milman, Y. V., 1993,
Microindentation on W and Mo Oriented Single Crystals: an STM Study,
Acta Metall. Mater., 41, pp. 28552865.
10 Fleck, N. A., Muller, G. M., Ashby, M. F., and Hutchinson, J. W., 1994,
Strain Gradient Plasticity: Theory and Experiments, Acta Metall. Mater.,
422, pp. 475487.
11 Stolken, J. S., and Evans, A. G., 1998, A Microbend Test Method for Measuring the Plasticity Length Scale, Acta Mater., 4614, pp. 51095115.
12 Dinesh, D., Swaminathan, S., Chandrasekar, S., and Farris, T. N., 2001, An
Intrinsic Size-Effect in Machining Due to the Strain Gradient, Proc. 2001
ASME IMECE, Nov. 1116, New York, pp. 197204.
13 Joshi, S. S., and Melkote, S. N., 2004, An Explanation for the Size-Effect in
Machining Based on Strain Gradient Plasticity, ASME J. Manuf. Sci. Eng.,
1264, pp. 679684.
14 Liu, K., and Melkote, S. N., 2004, A Strain Gradient Based Finite Element
Model for Micro/Meso-Scale Orthogonal Cutting Process, Proceedings of
2004 Japan-USA Symposium on Flexible Automation, Denver, CO, July.
15 Nakayama, K., and Tamura, K., 1968, Size Effect in Metal-Cutting Force,
American Society of Mechanical Engineers, 67-Prod-9, 1967.
16 Kim, K. W., Lee, W. Y., and Sin, H. C., 1999, A Finite Element Analysis of
Machining With the Tool Edge Considered, J. Mater. Process. Technol., 86,
pp. 4555.
17 Lucca, D. A., Rhorer, R. L., and Komanduri, R., 1993, Effect of Tool Edge
Geometry on Energy Dissipation in Ultraprecision Machining, CIRP Ann.,
421, pp. 8386.
18 Armarego, E. J. A., and Brown, R. H., 1962, On the Size Effect in Metal
Cutting, Int. J. Prod. Res., 13, pp. 7599.
19 Komanduri, R., Chandrasekaran, N., and Raff, L. M., 1998, Effect of Tool
Geometry in Nanometric Cutting: a Molecular Dynamics Simulation Approach, Wear, 219, pp. 8497.
20 Atkins, A. G., 2003, Modelling Metal Cutting Using Modern Ductile Fracture
Mechanics: Quantitative Explanations for Some Longstanding Problems, Int.
J. Mech. Sci., 45, pp. 373396.
21 Gao, H., and Huang, Y., 2001, Taylor-Based Nonlocal Theory of Plasticity,
Int. J. Solids Struct., 38, pp. 26152637.
22 Johnson, G. R., and Cook, W. H., 1983, A Constitutive Model and Data for
Metals Subjected to Large Strain, High Strain Rates and High Temperatures,
Proc. 7th Int. Symposium on Ballistics, April, The Hague, Netherlands, pp.
541547.
23 Liu, K., 2005, Process Modeling of Micro-Cutting Including Strain Gradient
Effect, Ph.D thesis, Mechanical Engineering, Georgia Institute of Technology.
24 Fleck, N. A., and Hutchinson, J. W., 1993, A Phenomenological Theory for
Strain Gradient Effects in Plasticity, J. Mech. Phys. Solids, 41, pp. 1825
1857.
25 Gao, H., Huang, Y., Nix, W. D., and Hutchinson, J. W., 1999, Mechanism-

738 / Vol. 128, AUGUST 2006

26
27
28

29
30
31

32
33
34
35
36
37
38

39
40
41
42

Based Strain Gradient PlasticityI. Theory, J. Mech. Phys. Solids, 47, pp.
12391263.
Acharya, A., and Bassani, J. L., 2000, Lattice Incompatibility and a Gradient
Theory of Crystal Plasticity, J. Mech. Phys. Solids, 48, pp. 15651595.
Zorev, N. N., 1963, Inter-Relationship Between Shear Processes Occurring
Along Tool Face and Shear Plane in Metal Cutting, International Research in
Production Engineering, pp. 4249.
Usui, E., and Shirakashi, T., 1982, Mechanics of MachiningFrom Descriptive to Predictive Theory. In on the Art of Cutting Metals75 Years Later,
Proceedings of International Conference on Production Engineering Research,
September 912, Pittsburgh, PA, ASME, New York, PED-7, pp. 1335.
Shih, A. J., Chandrasekar, S., and Yang, H. T., 1990, Finite Element Simulation of Metal Cutting Process With Strain-Rate and Temperatures Effects,
Fundamental Issues in Machining, ASME, PED, 43, pp. 1124.
Bhattacharya, S., and Lovell, M. R., 2000, Characterization of Friction in
Machining: Evaluation of Asperity Deformation and Seizure-Based Models,
Trans. NAMRC/SME, XXVII, pp. 107112.
Lovell, M. R., Bhattacharya, S., and Zeng, R., 1998, Modeling Orthogonal
Machining for Variable Tool-Chip Interfacial Friction Using Explicit Dynamic
Finite Element Methods, Proceedings of the CIRP International Workshop on
Modeling of Machining Operations, Atlanta, GA, pp. 265276.
Tao, Z., and Lovell, M. R., 2002, Towards an Improved Friction Model in
Material Removal Processes: Investigating the Role of Temperature, SME
Technical Paper No. MR02-188, pp. 18.
Shi, G., Deng, X., and Shet, C., 2002, A Finite Element Study of the Effect of
Friction in Orthogonal Metal Cutting, Finite Elem. Anal. Design, 38, pp.
863883.
Madhavan, V., Chandrasekar, S., and Farris, T. N., 2000, Machining as a
Wedge Indentation, ASME J. Appl. Mech., 67, pp. 128139.
Ceretti, E., Fallbohmer, P., Wu, W. T., and Altan, T., 1996, Application of 2D
FEM to Chip Formation in Orthogonal Cutting, J. Mater. Process. Technol.,
59, pp. 169180.
Hamel, V., Roelandt, J. M., Gacel, J. N., and Schmidt, F., 2000, Finite Element Modeling of Clinch Forming With Automatic Remeshing, Comput.
Struct., 77, pp. 185200.
Fang, N., 2003, Sensitivity Analysis of the Material Flow Stress in Machining, American Society of Mechanical Engineers, Manufacturing Engineering
Division, MED, 14, pp. 2332.
Clausen, A. H., Borvik, T., Hopperstad, O. S., and Benallal, A., 2004, Flow
and Fracture Characteristic of Aluminum Alloy AA5083-H116 as Function of
Strain Rate, Temperature and Triaxiality, Mater. Sci. Eng., A, 364, pp. 260
272.
Altan, T., and Bougler, F. W., 1973, Flow Stress of Metals and its Application
in Metal Forming Analyses, American Society of Mechanical Engineers, No.
73, Prod-4.
Oxley, P. L. B., 1989, Mechanics of Machining: An Analytical Approach to
Assessing Machinability, Ellis Horwood Ltd., Chichester.
Zhou, M., and Clode, M. P., 1998, Constitutive Equations for Modeling Flow
Softening Due to Dynamic Recovery and Heat Generation During Plastic Deformation, Mech. Mater., 27, pp. 6376.
MatWeb material database http://www.matweb.com.

Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 07/17/2015 Terms of Use: http://asme.org/terms

You might also like