You are on page 1of 15

Int. J. Machining and Machinability of Materials, Vol. 6, Nos.

1/2, 2009 159

Enhanced finite element model of ultrasonically


assisted turning

Naseer Ahmed
Ghulam Ishaq Khan Institute of Engineering
Sciences and Technology,
Topi, Swabi, N.W.F.P., Pakistan
Fax: +92-938-271889
E-mail: naseer@giki.edu.pk

Alexander V. Mitrofanov,* Vladimir I. Babitsky


and Vadim Silberschmidt
Wolfson School of Mechanical and Manufacturing Engineering,
Loughborough University,
Loughborough, Leicestershire LE11 3TU, UK
Fax: +44-1509-227648
Fax: +44-1509-227502
E-mail: a.v.mitrofanov@lboro.ac.uk
E-mail: v.i.babitsky@lboro.ac.uk
E-mail: v.silberschmidt@lboro.ac.uk
*Corresponding author

Abstract: This paper presents a new enhanced computational model of


Ultrasonically Assisted Turning (UAT) that is based on the DEFORM 3D FE
code and allows transient, coupled thermo-mechanical simulations of both
UAT and Conventional Turning (CT) of elasto-plastic materials. The model
features 3D oblique chip formation and realistic geometry of a cutting tool. The
model is used to analyse the effect of cutting and vibration parameters
on cutting forces in UAT. The comparison between efficiency of various
vibration directions is also carried out. Optimum values for cutting and
vibration parameters are suggested as a result of this numerical study.

Keywords: finite element modelling; turning; ultrasonic vibration.

Reference to this paper should be made as follows: Ahmed, N.,


Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V. (2009) ‘Enhanced finite
element model of ultrasonically assisted turning’, Int. J. Machining and
Machinability of Materials, Vol. 6, Nos. 1/2, pp.159–173.

Biographical notes: Naseer Ahmed is an Assistant Professor at Ghulam Ishaq


Khan Institute of Engineering Sciences and Technology in Pakistan. He earned
his PhD in Mechanical Engineering from Loughborough University, UK. His
research interests include finite element modelling of machining processes and
ultrasonically assisted turning. He has published in journals such as the Journal
of Materials Processing Technology, Computational Materials Science and
Journal of Sound and Vibration.

Copyright © 2009 Inderscience Enterprises Ltd.


160 N. Ahmed et al.

Alexander V. Mitrofanov is a Lecturer in Mechanics of Materials and a


Member of Mechanics of Advanced Materials Research Group. Educated at
St. Petersburg State Technical University, Russia, he obtained his PhD in
Mechanical Engineering from Loughborough University. He teaches
undergraduate and MSc courses in Mechanics of Materials, Finite Element
Analysis and Advanced Manufacturing Technology. His research areas of
interest include computational and experimental analyses of material
processing techniques, ultrasonically assisted machining and advanced
materials, for example, superalloys and biomaterials.

Vladimir I. Babitsky is a Professor of Dynamics at Loughborough University,


UK and a Head of Non-linear Dynamics Research Group. He graduated from
Moscow State Technological University and received his PhD and DSc in
Non-linear Dynamics from Russian Academy of Sciences. He is the author of
7 monographs and handbooks in non-linear dynamics, vibration control and
machine dynamics, more than 200 journal and conference papers and more
then 40 Russian and international patents. He is an organiser and current
president of the International Centre of Vibro-Impact Systems, and an
originator and editor of Springer book series ‘Foundations of Engineering
Mechanics’.

Vadim Silberschmidt is a Professor in Mechanics of Materials at


Loughborough University, UK and the Head of the Mechanics of Advanced
Materials Research Group. He obtained his PhD from Russian Academy of
Sciences, Perm. He is a Fellow and the Director of the International Centre of
Vibro-Impact Systems. He teaches postgraduate modules in Structural Integrity
and Finite Element Analysis. His research interests include fracture and
continuum damage mechanics, computational material sciences and
micromechanics of materials. He has published in journals such as
International Journal of Fracture, Engineering Fracture Mechanics and
Materials Science and Engineering A.

1 Introduction

Ultrasonically Assisted Turning (UAT) is an advanced machining technique where


high-frequency vibration is superimposed on the movement of the cutting tool.
Compared to Conventional Turning (CT), UAT leads to a higher cutting performance,
yielding a significant decrease in cutting forces and superior surface finish. UAT also
enables machining of intractable materials, such as high-strength aerospace alloys,
composites and brittle ceramics. Though being of such significance, this technique has
not yet been widely introduced in the industry. Problems such as instability of the cutting
process that resulted in poor finish prevented the full implementation of this process.
The prototype of the UAT system has been designed at Loughborough University,
UK, and various experimental tests were carried out to study the advantages and
shortcomings of UAT in comparison to CT. There have been several attempts to improve
UAT by introduction of advanced autoresonant control (Babitsky et al., 2004a,b) and by
studying the dynamics of UAT as a non-linear vibro-impact process the cutting process
(Astashev and Babitsky, 1998). These attempts allowed us to enhance the cutting
performance of UAT, however, there was a lack of understanding of mechanics of the
tool–workpiece interaction in UAT and of the influence of various cutting and vibration
Enhanced finite element model of UAT 161

parameters on the vibration-assisted cutting process. Since analytical approaches cannot


be applied to solve this problem due to its complex geometry and non-linear material
behaviour, numerical simulations were further employed.
Finite Element Method (FEM) is a main computational tool used to model the
process zone and tool–workpiece interaction in metal cutting. This technique had
been applied to the cutting simulation for nearly four decades now. A predefined
separation line together with an ‘unzipping’ algorithm were used to simulate the chip
formation in Carroll and Strenkowski (1988), Lin and Pan (1993), Shih (1995) and
Dokainish et al. (1988), while continuous remeshing–rezoning not requiring a separation
criterion was used to model the chip separation in Shinozuka et al. (1996), Ng et al.
(1999), Monaghan and MacGinley (1999) and Ozel and Altan (2000b). 3D modelling of
machining processes was also performed by many researchers using both the predefined
area for chip separation (Lin and Lin, 1999; Lin and Yarng, 1997; Soo et al., 2004;
Strenkowski et al., 2002) and the remeshing–rezoning technique (Ahmed et al., 2006;
Fang and Zeng, 2005; Mitrofanov et al., 2005; Pantale et al., 2004). Finite element
simulations were used to model UAT for the first time in Mitrofanov et al. (2003),
Babitsky et al. (2004) and Ahmed et al. (2006).
The FE model of UAT was initially developed with MSC.MARC/MENTAT general
FE code and presented at the 8th CIRP Workshop (Ahmed et al., 2005). The paper
discussed both 2D and 3D formulation of the UAT process in comparison with the CT
one, and presented results on the chip formation process, cutting forces, effect of
lubrication friction and thermal processes in turning. That model was further used for the
analysis of burr formation in drilling with and without ultrasonic vibration in Stoll et al.
(2006). A detailed description of the MARC.MENTAT model for UAT and other key
findings can also be found in Mitrofanov et al. (2005), Ahmed et al. (2006) and Ahmed
et al. (2007).
Although, the FE code MARC.MENTAT proved to be helpful to perform
simulations of orthogonal cutting, it is not capable to solve problems with the real
geometry of oblique cutting. The main reason for this is that MARC.MENTAT demands
a high volume of memory and higher processing power, which currently are not
available. Hence, an advanced specialised FE code for metal cutting had to be employed.
DEFORMTM 3D (2005) is a special purpose FE software for large-deformation cases that
is capable to model complex 3D material flow patterns. The advantages of DEFORMTM
are a fully automatic mesh generator with a local control of the element size, easy-to-use
templates for quicker modelling and support for user-defined subroutines for material
modelling and fracture criteria. DEFORMTM uses Arbitrary Langrangian-Eulerian (ALE)
method based on the idea of separating the material and mesh displacements to
eliminate mesh distortion as compared to the Updated Langrangian method used in
MARC.MENTAT. The mesh refinement scheme, adopted in ALE, works effectively and
efficiently. In the last several years, DEFORMTM had won the attention of many
researchers working in the area of simulations of machining operations (see e.g. Bil
et al., 2004; Ozel and Altan, 2000a).
Applying DEFORMTM to the simulations of UAT did not prove to be a
straightforward procedure and a considerable effort was put into overcoming the
limitations of the software at the model development phase. Section 2 explains the
development of the FE model of UAT in DEFORMTM.
162 N. Ahmed et al.

2 Model description

The material properties in the model developed with DEFORMTM are the same as those
used in MARC.MENTAT models (Ahmed et al., 2005): a modified Johnson–Cook
material model is used with maximum stresses capped at the ultimate tensile strength
value for Inconel 718. The advantage offered by DEFORMTM is the modelling of a full
tool rather than modelling only the tool tip as was the case previously. Simulations with
DEFORMTM did not use the RAM of the computer directly but rather were relaying on
the processing power, thus providing considerable computational benefits in comparison
with MARC.MENTAT. This resulted in shorter simulation times: a successful
completion of one simulation with MARC.MENTAT required about 7 hr calculation
time on Pentium IV, 2.8 GHz workstation with 2048 MB RAM, whereas a full run for
the model in DEFORMTM required only 2 hr on the same workstation.
Modelling of the real geometry improves a chance to predict forces, stresses and
temperature close to experimental values in this type of simulation. Figure 1 shows the
area of the workpiece–tool interaction that was considered in the FE simulation.
Both straight and curved workpiece surfaces were modelled in the simulations, with the
radius of curvature corresponding to that of the workpiece circular cross section in the
latter case.

Figure 1 An analysis domain of the simulation model

As in the case of MARC.MENTAT-based simulations the model is fully


thermo-mechanically coupled. This approach also employs an ALE solver. The length of
the modelled workpiece is 2 mm. The ambient temperature is 20°C while the cutting tool
has the initial temperature 70°C – to reflect the additional heating due to ultrasonic
vibration of the tool. DNMA432 tool insert is selected from the library of available tools
of the software with the tool’s nose radius being 0.79 mm. The cutting tool is discretised
into about 12,000 four-noded tetrahedral elements with a mesh density being higher at
the tool tip (Figure 2(a)). The contact heat transfer coefficient between the tool and
workpiece is 50 N/s/mm/°C. The workpiece is modelled as elasto-plastic; its initial
temperature is 33°C. It is meshed using about 40,000 four-noded tetrahedral elements
with a minimum element length 0.05 mm (Figure 2(b)). The software is automatically
controlling the mesh size in different areas and is keeping a higher mesh density in the
chip formation zone and around it, thus improving the simulation time by focusing more
on the tool–workpiece interaction zone, which is the main area of study. The number of
simulation steps is kept at 960 that cover 6 ms of the cutting time that is equal to
120 cycles of vibration with f = 20 kHz.
Enhanced finite element model of UAT 163

Figure 2 Meshed models of cutting tool (a) and workpiece (b) (see online version
for colours)

Different cases were simulated for UAT aiming at the comparison of the effects of the
cutting speed, friction condition, feed rate, direction of vibration, vibration amplitude
and vibration frequency. Various simulation parameters studied in this paper are shown
in Table 1. Models of CT were also simulated for comparison purposes. All the basic
simulation variables such as material properties and a friction model were the same
for the cases of CT and UAT, with the only difference being the absence of vibration in
case of CT.

Table 1 Variables of turning

Parameters Magnitudes used in FEA


Cutting speed (m/min) 10; 20; 30
Friction coefficient 0; 0.5
Feed rate (mm/rev) 0.1; 0.2; 0.3
Vibration amplitude (µm) 7.5; 15; 30
Frequency (kHz) 10; 20; 30

Unlike the MARC.MENTAT model formulation (Ahmed et al., 2006, 2007; Mitrofanov
et al., 2005), the kinematic boundary conditions had to be applied only to the tool, that is,
the tool was moved with a constant velocity along the X-axis (Figure 1) with ultrasonic
vibration superimposed onto this a continuous motion (see Figure 3). This was due to the
software limitation that continuous movement of the workpiece could be ‘lost’ after the
remeshing procedure. The following equation was used to calculate the final boundary
conditions for the cutting tool having both vibration and constant velocity:
⎛ Lt ⎞
d (t ) = ⎜ + a sin(2π ft ) ⎟ (1)
⎝ τ ⎠

Figure 3 Displacement of cutting tool


164 N. Ahmed et al.

where d(t) is the position of the cutting tool; a and f are the amplitude and frequency of
vibration, respectively, L is the total length of cut, t is the instantaneous time and τ is the
total simulation time.

3 Simulation results and discussion

Figure 4 shows an example of the FE simulation presenting the temperature distribution


in the tool–workpiece interaction zone. Simulations performed on the basis of the
variables presented in Table 1 resulted in a large number of cases and important results.
For the ease of understanding, the results have been studied separately focusing on the
impact of a unique parameter.

Figure 4 Temperature distributions in cutting regions for UAT (see online version
for colours)

3.1 Effect of vibration direction


The effect of the vibration direction on the overall performance of the turning process
was analysed. Three different cases – CT model, UAT model with vibration in the
tangential direction (i.e. along X-axis in Figure 1) and UAT model with vibration in the
feed direction (i.e. along the Z-axis) – were analysed. Figure 5 shows the comparison of
cutting forces acting on the cutting tool for those three cases.

Figure 5 Comparison of forces for three different types of turning modes


Enhanced finite element model of UAT 165

The forces in case of CT have a practically constant magnitude, as the tool is always
interacting with the workpiece. In case of UAT, with vibration applied in the feed
direction, the forces also have a nearly constant magnitude. Although the cutting tool
vibrates in the feed direction, it still remains in contact with the freshly formed chip tool,
and that is the reason for such a character of the force magnitude.
The forces in case of UAT with vibration in the tangential direction are close to zero
when the tool is not in contact with the chip, but once the tool comes in contact with the
chip, forces rise sharply. When the tool starts to move away from the separated chip at
the subsequent stage of the cycle of ultrasonic vibration, the forces rapidly decline back
to zero. The principal difference between application of vibration in the tangential and
feed directions is that in case of UAT in the tangential direction, in every vibration cycle,
the cutting tool moves along the direction of workpiece motion, while in case of UAT in
the feed direction, the cutting tool oscillates along a normal to the direction of workpiece
motion and thus maintains a contact with the chip throughout the cycle of vibration.
Disengagement between the tool and the chip in case of tangential vibrations results
in a force drop to zero. The difference in forces between three studied cases becomes
apparent when average forces for a single vibration cycle are compared (Figure 6).
A force drop of 46% is observed for transformation of the turning mode from CT to
UAT with vibration in the feed direction and a further drop of 30% is observed when
vibration is applied in the tangential direction.

Figure 6 Average forces for three types of turning

3.2 Effect of vibration amplitude


The impact of increasing amplitude on the overall performance of UAT was studied.
Figure 7 shows the difference in forces resulting from changing the amplitude value from
7.5 to 30 µm. An increase in the vibration amplitude caused an increase in the maximum
force within a complete cycle of vibration. Although from Figure 7, the amplitude
7.5 µm seems to lower forces, examination of their average magnitudes for ultrasonic
vibration (Figure 8) shows that the cutting force for amplitude 7.5 µm is in fact higher
than that in cases of 15 and 30 µm. A 16% drop in the force magnitude results when the
amplitude is increased from 7.5 to 15 µm, while a further increase in the amplitude from
15 to 30 µm does not cause any significant change in the average force. It indicates that
amplitude has an impact up to a certain level, after which a further increase does not
affect the forces. It is important to mention here that this change in forces was studied for
a cutting speed of 20 m/min. A different optimum value of vibration amplitude can exist
for a different cutting speed.
166 N. Ahmed et al.

Figure 7 Effect of ultrasonic amplitude on cutting forces

Figure 8 Forces averaged for one complete ultrasonic vibration cycle: effect of
vibration amplitude

3.3 Effect of vibration frequency


The effect of vibration frequency on the overall performance of UAT was also studied.
The peak values of forces (Figure 9) for different frequency values remain the same,
as the amplitude is constant and thus the tool penetration depth into the workpiece per
cycle of vibration does not change. But examining the average values of forces
(Figure 10) for the effect of frequency shows that as the frequency increases the time
average of forces decreases. Interestingly, there exist an upper limit for the maximum
beneficial value of frequency, beyond which the forces start increasing. Here, the average
over time is necessary to account for different duration of cycles at different frequencies.

Figure 9 Effect of ultrasonic frequency on cutting forces


Enhanced finite element model of UAT 167

Figure 10 shows that the increasing frequency from 10 to 20 kHz results in an average
force drop of 35%. A further drop of 38% is observed when the frequency value is
increased from 20 to 30 kHz. However, this trend changes with the next increase of
frequency from 30 to 40 kHz as the average force value in fact increases by about 10%.
This shows that frequency has an optimum magnitude for beneficial utilisation of UAT.

Figure 10 Effect of frequency: time average of forces

3.4 Effect of feed rate


The effect of increasing feed rate in turning on the overall forces in the cutting tool was
analysed. Three different feed rates of 0.1, 0.2 and 0.3 mm/rev were examined for UAT
at a fixed cutting speed of 20 m/min and friction level μ = 0.5 . Figure 11 shows the
comparison of forces for these three cases. The maximum force values resulted as
116, 261 and 568 N for feed rates of 0.1, 0.2 and 0.3 mm/rev, respectively.

Figure 11 Effect of feed rate on cutting forces

Examining the average values of forces for three feed rates over one complete vibration
cycle in Figure 12 revealed a 127% increase in the force when the feed rate was
increased from 0.1 to 0.2 mm/rev and a further increase of approx. 250% (a 110%
increase from force value at feed rate 0.2 mm) when the feed rate was increased from
0.2 to 0.3 mm/rev.
168 N. Ahmed et al.

Figure 12 Effect of feed rate: average forces

From examination of this increase, it can be concluded that feed rate of 0.2 mm/rev is an
optimum value for UAT with a cutting speed of 20 m/min and μ = 0.5 . Increase in the
feed rate from 0.2 to 0.3 mm/rev results in double forces but only 50% more material
removal. Higher forces are detrimental for the life of a cutting tool insert, especially,
when machining high-strength alloys.
The effect of the feed rate on cutting forces was also studied by Hua et al. (2005).
The findings suggested that more deformation energy was required when the feed rate
was increased. The feed rate influenced the cutting forces more than any other variable,
and cutting forces increased by same factor (110%) for an increase in the feed rate from
0.14 to 0.28 mm/rev in hard turning of bearing steel.

3.5 Effect of friction


Simulations were conducted to analyse the impact of the coefficient of friction on the
overall cutting forces in the cutting tool. Two different coefficients of friction were
introduced, that is, μ = 0 and μ = 0.5 . Figure 13 shows the forces in the cutting tool for
two cases. At a feed rate 0.2 mm/rev, a big difference of forces due to friction is
observed between these cases. The force increases by 67% when the coefficient of
friction is increased from 0 to 0.5. This difference varies for different feed rates, as can
be seen in Figure 14, when the feed rate is equal to 0.3 mm/rev. In this case, a 33%
increase in forces is observed when coefficient of friction is increased from 0 to 0.5.

Figure 13 Effect of coefficient of friction on cutting forces; feed rate 0.2 mm/rev
Enhanced finite element model of UAT 169

Figure 14 Effect of coefficient of friction on cutting forces; feed rate 0.3 mm/rev

3.6 Comparison of forces for different cutting speeds


Simulations were also performed to study the effect of a changing cutting speed on the
forces in the cutting tool. Three different cutting speeds – 10, 20 and 30 m/min – were
considered at a constant feed rate 0.2 mm/rev and a coefficient of friction μ = 0.5 .
The results obtained are presented in Figure 15 with forces in the cutting tool attaining
the highest value for a maximum cutting speed.

Figure 15 Effect of cutting speed on forces

Figure 16 shows the average forces in the three cases for a single vibration cycle and
reveals that a 76% increase of forces occurs when the cutting speed is increased from
10 to 20 m/min and an increase of 33% in cutting forces occurs for an increase in the
cutting speed from 20 to 30 m/min.

Figure 16 Average forces for a single ultrasonic vibration cycle


170 N. Ahmed et al.

An increase in the cutting forces with an increase in the cutting speed was also observed
by Korkut and Donertas (2007) while machining AISI 1020 and AISI 1040 steels.
A tendency for high built-up edge during machining at low and intermediate cutting
speeds was considered to be a factor behind the increase of cutting forces with increase
in cutting speeds.
Also Mitrofanov (2004) experimentally observed a decrease in forces for CT from
140 to 130 N for an increase in cutting speeds from 40 to 125 rev/min, whereas an
increase in forces from 60 to 90 N was observed for UAT. A drop in the CT force was
explained by the lower fracture toughness of the workpiece material at higher strain rates
resulted from greater cutting speeds, while the increase in the UAT force was caused by
the increased time of the contact between the cutting tool and chip, with the cutting speed
getting closer to its critical value (ν t = 2π af ).

4 Conclusions

Three-dimensional thermo-mechanical finite element models of both UAT and CT


were investigated in this paper to study the effect of vibration and cutting parameters
on the performance of these cutting techniques. The presented DEFORM model
features a range of new capabilities as compared to the previous MARC.MENTAT
model presented at the 8th CIRP Workshop, such as the real geometry of the cutting
tool, oblique chip formation, a curved workpiece surface and localised mesh
refinement in the process zone. The current model was advantageous in many regards
if compared to both 2D models and the simplified 3D models (MARC.METAT).
The 2D model was only capable of studying orthogonal turning still lacking the
through-thickness-dimensions. The simplified 3D model (Ahmed et al., 2007) was
capable of simulation both orthogonal and oblique turning, but still the tool insert
geometry was different from that of the real tool. The 3D model presented in this paper
had thus overcome many if not all the shortcomings of the previous attempts because of
its capability to represent a more realistic turning model. A reduced workpiece and tool
insert size was considered in order to reduce simulation times that enabled consideration
of a significant number of case studies allowing selection of optimum process
parameters.
Simulations demonstrated considerable differences in cutting forces between CT and
UAT: reductions in cutting forces with introduction of ultrasonic vibration of the tool
was observed. A comparison of cases with various vibration directions was performed
and analysis suggested that tangential vibration in UAT leads to the highest force
reduction.
Conducted simulations were helpful in understanding the impact of changing
vibrations parameters, namely amplitude and frequency, on the overall forces in the
cutting tool. Both the amplitude and frequency were found to have their optimum values
for a certain cutting speed. Changing this speed will require determination of a different
optimum value for vibration frequency and amplitude. As shown by simulations,
increasing the amplitude from 7.5 to 15 µm resulted in a drop of forces for a cutting
speed of 20 m/min, but a further increase in the amplitude from 15 to 30 µm does not
affect the forces. The same trend was true for frequency, which also was discovered to
have an upper limit for the optimum performance at a certain speed. At 20 m/min,
Enhanced finite element model of UAT 171

increasing the frequency from 10 to 20 kHz and then to 30 kHz, resulted in a


reduced cutting force. But a further frequency increase from 30 to 40 kHz led to a force
growth by 10%.
The numerical study was also helpful to predict forces in the cutting tool for various
feed rates. Three different feed rates 0.1, 0.2 and 0.3 mm/rev were considered. Increasing
the feed rate from 0.1 to 0.2 mm/rev resulted in a nearly 127% increase in the force
acting on the cutting tool, whereas a further increase in the feed rate from 0.2 to
0.3 mm/rev resulted in a 110% increase in forces. Feed rate 0.2 mm/rev was therefore
recommended as an optimum value at a cutting speed of 20 m/min in order to extend the
tool life in machining of high-strength materials.
The effect of friction was also studied in order to estimate the influence of lubrication
on UAT. Friction had a different effect on forces for various feed rates. At feed rate
0.2 mm/rev and cutting speed 20 m/min, changing friction coefficient from 0 to
0.5 resulted in a 67% increase in forces, while at the same cutting speed but with feed
rate 0.3 mm/rev, only a 33% increase was obtained. Although simulations predict a drop
in cutting forces when lubrication is used, in practice the lubricant does not fully
penetrate into the tool–workpiece interaction zone due to high contact pressures in
turning. But due to the non-regular nature of contact in case of UAT, an improved access
of the lubricant into the tool–workpiece interaction is provided. Mitrofanov (2004)
experimentally observed a decrease in forces up to 45% in case of UAT when lubrication
was introduced.
Simulations also allowed estimations of cutting forces for different cutting speeds.
In case of UAT cutting forces increase with an increase in the cutting speed from
10 to 30 m/min. An increase of 76% was observed when the cutting speed was increased
from 10 to 20 m/min. But when the cutting speed was increased from 20 to 30 m/min,
the increase was only 33%. A non-permanent contact is again the reason for the force
reduction in UAT, but as the cutting speed is increased, the contact time between the tool
and workpiece increased, which resulted in growing cutting forces.

References
Ahmed, N., Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V.V. (2005) ‘Finite element
modeling of ultrasonically assisted turning’, 8th CIRP International Workshop on Modeling of
Machining Operations, Chemnitz, Germany, pp.107–114.
Ahmed, N., Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V.V. (2006a) ‘3D finite element
analysis of ultrasonically assisted turning’, Computational Materials Science, Vol. 39,
pp.149–154.
Ahmed, N., Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V.V. (2006b) ‘Analysis of
material response to ultrasonic vibration loading in turning Inconel 718’, Material Science and
Engineering A, Vol. 424, pp.318–325.
Ahmed, N., Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V.V. (2007a) ‘Analysis of forces
in ultrasonically assisted turning’, Journal of Sound and Vibration, available online,
doi:10.1016/j.jsv.2007.04.003.
Ahmed, N., Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V. (2007b) ‘3D finite element
analysis of ultrasonically assisted turning’, Computational Materials Science,
Vol. 39, pp.149–154.
Astashev, V.K. and Babitsky, V.I. (1998) ‘Ultrasonic cutting as a nonlinear (vibro-impact)
process’, Ultrasonics, Vol. 36, pp.89–96.
172 N. Ahmed et al.

Babitsky, V.I., Kalashnikov, A. and Astashev, V.K. (2004) ‘Autoresonant control of nonlinear
mode in ultrasonic transducer for machining applications’, Ultrasonics, Vol. 42, pp.29–35.
Babitsky, V.I., Kalashnikov, A.N. and Molodtsov, F.V. (2004) ‘Autoresonant control of
ultrasonically assisted cutting’, Mechatronics, Vol. 14, pp.91–114.
Babitsky, V.I., Mitrofanov, A.V. and Silberschmidt, V.V. (2004) ‘Ultrasonically assisted turning of
aviation materials: simulations and experimental study’, Ultrasonics, Vol. 42, pp.81–86.
Bil, H., Kilic, S.E. and Tekkaya, A.E. (2004) ‘A comparison of orthogonal cutting data from
experiments with three different finite element models’, International Journal of Machine
Tools and Manufacture, Vol. 44, pp.933–944.
Carroll, J.I. and Strenkowski, J. (1988) ‘Finite, element models of orthogonal cutting with
application to single point diamond turning’, International Journal of Mechanical Science,
Vol. 30, pp.899–920.
DEFORMTM 3D Version 5.1 User's Manual (2005) Scientific Forming Technologies Corporation,
Columbus, OH.
Dokainish, M.A., Elbestawi, M.A., Polat, U. and Tole, B. (1988) ‘Analysis of stresses during exit in
interrupted cutting with chamfered tools’, International Journal of Machine Tools and
Manufacture, Vol. 29, pp.519–534.
Fang, G. and Zeng, P. (2005) ‘Three-dimensional thermo-elastic-plastic coupled FEM simulations
for metal oblique cutting process’, Journal of Materials Processing Technology, Vol. 168,
pp.42–48.
Hua, J., Shivpuri, R., Cheng, X., Bedekar, V., Matsumoto, Y., Hashimoto, F. and
Watkins, T.R. (2005) ‘Effect of feed rate, workpiece hardness and cutting edge on subsurface
residual stress in the hard turning of bearing steel using chamfer + hone cutting edge
geometry’, Material Science and Engineering A, Vol. 394, pp.238–248.
Korkut, I. and Donertas, M.A. (2007) ‘The influence of feed rate and cutting speed on the cutting
forces, surface roughness and tool-chip contact length during face milling’, Materials and
Design, Vol. 28, pp.308–312.
Lin, Z-C. and Lin, Y-Y. (1999) ‘A study of an oblique cutting modelling’, Journal of Materials
Processing Technology, Vol. 86, pp.119–130.
Lin, Z.C. and Pan, W.C. (1993) ‘A thermoelastic-plastic large deformation model for orthogonal
cutting with tool flank wear-part I: computational procedures’, International Journal of
Mechanical Science, Vol. 35, pp.829–840.
Lin, Z.C. and Yarng, Y.D. (1997) ‘Three-dimensional cutting process analysis with different
cutting velocities’, Journal of Materials Processing Technology, Vol. 70, pp.22–33.
Mitrofanov, A.V. (2004) Modelling the Ultrasonically Assisted Turning of High-Strength Alloys,
Loughborough University, Wolfson School of Mechanical and Manufacturing Engineering.
Mitrofanov, A.V., Ahmed, N., Babitsky, V.I. and Silberschmidt, V.V. (2005) ‘Effect of lubrication
and cutting parameters on ultrasonically assisted turning of Inconel 718’, Journal of Materials
Processing Technology, Vols. 162–163, pp.649–654.
Mitrofanov, A.V., Babitsky, V.I. and Silberschmidt, V.V. (2003) ‘Finite element simulations of
ultrasonically assisted turning’, Computational Materials Science, Vol. 28, pp.645–653.
Monaghan, J. and MacGinley, T. (1999) ‘Modelling the orthogonal machining process using coated
carbide cutting tools’, Computational Materials Science, Vol. 16, pp.275–284.
Ng, E.G., Aspinwall, D.K., Brazil, D. and Monaghan, J. (1999) ‘Modelling of temperature and
forces when orthogonally machining hardened steel’, International Journal of Machine Tools
and Manufacture, Vol. 39, pp.885–903.
Ozel, T. and Altan, T. (2000a) ‘Determination of workpiece flow stress and friction at the chip–tool
contact for high-speed cutting’, International Journal of Machine Tools and Manufacture,
Vol. 40, pp.133–152.
Ozel, T. and Altan, T. (2000b) ‘Process simulation using finite element method – prediction of
cutting forces, tool stresses and temperatures in high-speed flat end milling’, International
Journal of Machine Tools and Manufacture, Vol. 40, pp.713–738.
Enhanced finite element model of UAT 173

Pantale, O., Bacaria, J.L., Dalverny, O., Rakotomalala, R. and Caperaa, S. (2004) ‘2D and 3D
numerical models of metal cutting with damage effects’, Computational Methods for Applied
Mechanical Engineering, Vol. 193, pp.4383–4399.
Shih, A.J. (1995) ‘Finite element analysis of the rake angle effects in orthogonal metal cutting’,
International Journal of Mechanical Science, Vol. 38, pp.1–17.
Shinozuka, J., Obikawa, T. and Shirakashi, T. (1996) ‘Chip breaking analysis from the viewpoint of
the optimum cutting tool geometry design’, Journal of Materials Processing Technology,
Vol. 62, pp.345–351.
Soo, S.L., Aspinwall, D.K. and Dewes, R.C. (2004) ‘3D FE modelling of the cutting of Inconel
718’, Journal of Materials Processing Technology, Vol. 150, pp.116–123.
Stoll, A., Ahmed, N., Mitrofanov, A.V., Silberschmidt, V. and Leopold, J. (2006) ‘Influence of
ultrasonically assisted cutting on burr formation’, 9th CIRP International Workshop on
Modeling of Machining Operations, Bled, Slovenia, pp.457–464.
Strenkowski, J.S., Shih, A.J. and Lin, J.C. (2002) ‘An analytical finite element model for predicting
three-dimensional tool forces and chip flow’, International Journal of Machine Tools and
Manufacture, Vol. 42, pp.723–731.

You might also like