Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Machining Science and Technology

An International Journal

ISSN: 1091-0344 (Print) 1532-2483 (Online) Journal homepage: https://www.tandfonline.com/loi/lmst20

Enhancing the machining performance by cutting


tool surface modifications: a focused review

Yuhan Chen, Jun Wang & Ming Chen

To cite this article: Yuhan Chen, Jun Wang & Ming Chen (2019) Enhancing the machining
performance by cutting tool surface modifications: a focused review, Machining Science and
Technology, 23:3, 477-509, DOI: 10.1080/10910344.2019.1575412

To link to this article: https://doi.org/10.1080/10910344.2019.1575412

Published online: 27 Mar 2019.

Submit your article to this journal

Article views: 502

View related articles

View Crossmark data

Citing articles: 17 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lmst20
MACHINING SCIENCE AND TECHNOLOGY
2019, VOL. 23, NO. 3, 477–509
https://doi.org/10.1080/10910344.2019.1575412

Enhancing the machining performance by cutting tool


surface modifications: a focused review
Yuhan Chena, Jun Wanga , and Ming Chenb
a
School of Mechanical and Manufacturing Engineering, UNSW Sydney (The University of New
South Wales), NSW 2052, Australia; bSchool of Mechanical Engineering, Shanghai Jiaotong
University, Shanghai, China

ABSTRACT KEYWORDS
In machining, cutting tools suffer from severe surface wear, Cutting tools; friction and
especially in the cutting of difficult-to-cut materials. A major wear; lubrication;
cause of tool wear is the friction generated at the tool-work machining; surface
modification;
and tool-chip interfaces, which produces a great deal of surface texturing
frictional heat and abrasion. In order to extend tool life and
improve the quality of machined components, a host of tech-
niques have been applied to modify the rake and flank faces
of cutting tools. These techniques aim at providing cutting
tools with improved resistance to external loading, better
tribological performance and/or better chemical stability. This
article presents a review of the fundamentals behind which
the friction and wear in machining are reduced by modifying
the cutting tool surface with the commonly used techniques,
such as surface coating, high energy beam treatment, and sur-
face texturing. The effects of these surface modifications on
improving the cutting performance are also analyzed. Future
research directions are finally discussed.

Introduction
Machining is an indispensable manufacturing process for various part
geometries for a wide range of engineering materials, such as metals,
ceramics, and composite materials. In machining, tool wear is a primary
concern which affects not only the productivity and costs, but also the
quality of the machined components. Tool wear takes places on both the
rake face and flank face of a cutting tool, and may be classified into four
major categories; namely abrasive wear, adhesive wear, diffusion wear, and
fatigue wear, although other forms of tool wear and damages have also
been reported (Childs, 2000). Friction at the tool-work and tool-chip inter-
faces, which is related to tool wear, is also a major concern as it affects the
cutting forces, heat generation and the occurrence of system vibration,
which in turn affects tool wear.

CONTACT Jun Wang jun.wang@unsw.edu.au School of Mechanical and Manufacturing Engineering,


UNSW Sydney (The University of New South Wales), NSW 2052, Australia.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/icnv.
ß 2019 Taylor & Francis Group, LLC
478 Y. CHEN ET AL.

To reduce friction and tool wear in machining, a large amount of


research effort has been made, including the development of advanced tool
materials (Song et al., 2012; Xu et al., 2013a), new and better cutting tool
designs (Wang and Zhang, 2008a, 2008b), optimization of processing param-
eters (Wang and Armarego, 2001; Wang et al., 2002), application of high-
performance cutting fluids (Yang et al., 2017; Wang et al., 2018), and cutting
tool surface modifications (Chen et al., 2002; Kawasegi et al., 2009). This
review has a particular interest on the cutting tool surface modifications.
The techniques for cutting tool surface modifications may be classified
into three categories, i.e. surface coating, surface treatment by high energy
beams, and surface texturing. Surface coating involves depositing a layer of
a material or multiple layers of different materials in several microns on
the tool surface. The coated layers of materials normally have high hard-
ness, low friction coefficients, stable chemical properties, and good thermal
isolating performance (Bouzakis et al., 2012). High energy beams, such as
electron or ion beams, have been used to treat the surfaces of cutting tools.
The irradiation of these energetic beams on the cutting tools can result in
major variations in the cutting tool microstructures and phases within the
surface layer, leading to an enhanced mechanical strength and tribological
performance. Surface texturing is a relatively new surface engineering tech-
nology, where micro-scale or nano-scale surface textures are fabricated on
the cutting tool through a variety of techniques. These textures are
designed to improve the tribological properties of cutting tool surfaces by,
for example, providing a better availability of lubricant at the tool-chip
interface, or reducing the tool-chip contact area. Additionally, surface tex-
tures can trap hard particles to reduce abrasion wear.
There have been published review works in the literature about surface
modification techniques and their effect on friction and wear in different
applications (Klocke and Krieg, 1999; Bouzakis et al., 2012; Aralan et al.,
2016; Gachot et al., 2017). Klocke and Krieg (1999) discussed the features
and applications of surface coated tools in metal machining, while Bouzakis
et al. (2012) discussed the surface coating technologies, the coating charac-
terization methods, and the effect of surface coating in cutting operations.
Gachot et al. (2017) reviewed the techniques of surface texturing for reduc-
ing friction and wear in general applications. In the work by Aralan et al.
(2016), surface texturing and its effect on cutting performance were dis-
cussed; however, two important phenomena induced by surface texturing
that play a crucial role in affecting the machining performance, i.e. the cut-
ting tool surface wettability and the actual tool-chip contact area, were
not discussed.
In this article, the common surface modification processes for cutting
tools are reviewed. Particular attention is paid to the mechanisms of the
MACHINING SCIENCE AND TECHNOLOGY 479

surface modifications on improving the cutting performance, especially tool


wear, in machining. Future research directions are also discussed.

Cutting tool surface coating


Cutting tool surface coatings are mainly achieved by two methods, chem-
ical vapor deposition (CVD) and physical vapor deposition (PVD). By
applying either a CVD or PVD technique, a thin layer of coating of several
microns with desired properties can be formed on the cutting tool surfaces.
In practice, multiple layers of coating of different materials may be made
on a cutting tool, where the inner layer close to the substrate has a better
affinity to the tool material while the outer layer has better properties for
cutting applications (Zeng et al., 2001).

Effect of hard coatings and relevant mechanisms


The first generation of hard coating materials for cutting tools is usually
known as hard carbon-based and nitride-based titanium compounds, such
as TiC, TiN, and TiCN. The initial applications of these coating materials
were basically monolayered. These coating materials have been very popu-
lar, mainly due to the following two factors. First, they usually combine a
wide range of good mechanical and thermal properties such as good hard-
ness and thermal conductivity; second, they can be deposited adequately
with good adhesion to the substrate (Klocke and Krieg, 1999). These fac-
tors have resulted in an improved cutting performance comparing to
uncoated tools. For example, Schintlmeister et al. (1984) found that the
cutting forces of TiC coated tools were 10% to 30% smaller than those of
uncoated HSS tools in the turning of a plain carbon steel; Lim et al. (1995)
reported that the flank wear rate of TiN coated HSS tools was much
lower than that of the uncoated tools in the dry turning of a medium-
carbon steel.
However, as machining processes are developing towards high-speed
and high-precision, those conventional hard coatings seem to be not good
enough for such severe cutting conditions, especially when more and
more difficult-to-cut materials are to be machined in industry, such as
high-speed milling of hardened steel (Klocke and Krieg, 1999). In high-
speed machining, cutting temperature is increased as cutting speed rises,
which leads to severe abrasion and adhesion at the tool-work and tool-
chip interfaces and increases the tendency for elemental diffusion and
oxidation. Although conventional coatings were still able to reduce tool
wear to some extent, their performance was not as good as expected in
these high-speed machining conditions. Furthermore, when machining
480 Y. CHEN ET AL.

some difficult-to-machine alloys, such as titanium alloys, the effect of


conventional coatings was also unsatisfactory. Klocke and Krieg (1999)
found that the tool life of TiN and TiCN coated HSS tools was even
shorter than that of uncoated HSS tools in the dry milling of Ti-6Al-4V
alloy. They further explained that the coated tools were more susceptible
to forming BUE due to the chemical affinity between the coating and the
work material.
The negative experiences with conventional hard coatings have encour-
aged a dramatic development of new cutting tool coatings. The main pur-
pose of such evolution was to further improve the cutting performance of
coated tools, especially for cutting under extremely tough cutting condi-
tions. The development includes new coating materials and new coating
structures. The evolution towards these two aspects and the resulting effect
on cutting performance (especially tool wear) are reviewed and discussed
respectively below.

Effect of improved coating materials


The development of hard coating materials started from adding other metal
elements into the titanium-based coatings to improve their mechanical,
thermal and chemical properties. A successful example is TiAlN. The add-
ition of aluminum is able to provide an additional strengthening effect on
the TiN crystal lattices and therefore increase the surface and subsurface
hardness, making it more resistant to abrasive wear and galling (Prengel
et al., 1997). Further, TiAlN has a good thermal isolating performance, due
to its low thermal conductivity (Leyendecker et al., 1991). Another advan-
tage of TiAlN coating is the possible formation of a protective Al2O3 layer
at the tool-chip interface during a cutting process, which provides the tool
surface with an increased chemical stability and oxidation resistance
(Leyendecker et al., 1991; Jindal et al., 1999). These improvements have
resulted in a better wear resistance than the conventional hard coatings.
For instance, Jindal et al. (1999) studied the cutting performance of TiAlN
coated tools in the turning of an Inconel 718 alloy, and showed that the
TiAlN coated tools exhibited a smaller flank wear rate than TiCN or
TiN coating.
Further improvement to the coating materials has been made by optimiz-
ing the content of Al in TiAlN. Generally, increasing the content of Al
gives a better thermal resistance but decreases the hardness of the coating
(Kalss et al., 2006), so that the Al content may be optimized according to
the process needs. In addition, improvements to the TiAlN properties have
been achieved by mixing some functional elements (typically Cr, Si, Y, and
Zr) to the TiAlN base, in an attempt to increase the wear resistance.
Compositions such as TiAlSiYN (Smyrnova et al., 2018) and TiAlZrN (Liu
MACHINING SCIENCE AND TECHNOLOGY 481

et al., 2016) are some of the latest good achievements of cutting tool
coating materials.
Ceramic coatings are another group of developed hard tool coatings.
In particular, alumina (Al2O3) has exhibited good wear resistance in
machining. Comparing to the conventional hard coatings, Al2O3 has a
higher hot-hardness, better chemical stability, and lower thermal conductiv-
ity. It also has an outstanding resistance to diffusion, owing to its dense
microstructure. In the study by M’saoubi and Ruppi (2009), a-Al2O3 coated
tools showed substantially less flank wear, smaller plastic deformation at
the cutting edge, and lower temperature within the tool substrate than
TiCN coated tools in turning an AISI 4140 steel.
A drawback of the alumina coatings is the possible phase change during
the cutting process. Alumina exists in many polymorphs. Among them,
a-Al2O3, c-Al2O3, and j-Al2O3 can be deposited in a controlled way on a
target surface. However, only a-Al2O3 is a stable phase, while c-Al2O3 and
j-Al2O3 are metastable and can transform to a-Al2O3 at high temperatures.
During such phase transformation, volume contraction can occur which
creates abundant porosity at the grain boundaries to reduce the adhesion
strength between the coating and the substrate (Ruppi, 2005), which may
cause the coating to peel off from the substrate and lead to tool failure.
Thus, the single-phase a-Al2O3 is more suitable for coating cutting tools.
A further group of cutting tool coating materials is known as superhard
coatings, among which CVD diamond has attracted the greatest attention.
This is not only due to its ultra-high hardness (ranging from 85  100 GPa
(Chowdhury et al., 2004)), but also many other excellent properties for cut-
ting applications. These properties include good thermal dissipation, small
thermal expansion, and superior chemical stability. Moreover, CVD dia-
mond has a small friction coefficient rarely exceeding l ¼ 0.2 in the air
(Field, 2012).
Although CVD diamond coated tools showed highly positive perform-
ance in most studies, such as in the drilling of SiC-reinforced aluminum
base composites (Chen et al., 2002) and in the high-speed end milling of
aluminum alloys (Torres et al., 2009), it has been found by Zhang et al.
(2010) that in the turning of a GFRP material, the wear of CVD diamond
coated WC-Co tools increased sharply after a certain cutting distance (or
cutting time), as shown in Figure 1. The sudden increase in the tool wear
rate, or in fact tool failure, was said to be a result of the peeling-off of the
coating from the substrate (Zhang et al., 2010). However, there has been
no study to enable a quantitative prediction of the occurrence of this phe-
nomena. Nevertheless, it is understood that this peeling-off is related to the
bonding strength between the diamond coating and the substrate which is
greatly affected by the deposition process and the surface condition of the
482 Y. CHEN ET AL.

Figure 1. Sudden failure of diamond coated tools at a certain cutting time in turning GFRP
materials (Zhang et al., 2010).

substrate (or cutting tool). The bonding strength should determine the cut-
ting time or distance at which the coating may peel off. It should be noted
that CVD diamond coating is not suitable for machining ferrous metals
due to the chemical affinity between C in diamond and Fe in the metals.
When examining the grain size of the coating materials, conventional
hard coating materials are usually composed of crystals with a grain size
ranging from several hundred nanometers to several hundred millimeters
(mono-crystals) (Musil, 2000). By reducing the grain size to the nanocrys-
talline range (typically around 10 nm) and bonding the nanocrystals using
another phase, a special group of costing materials with completely new
properties have been obtained, and those are known as nanocomposite
materials (Musil, 2000). As examples, hard nitrides (e.g. nc-TiN (Ma et al.,
2005), nc-TiAlN (Jilek et al., 2004), and nc-AlCrN (Liu et al., 2013)) have
been selected as the nanocrystalline phase (nc-) while amorphous Si3N4 (a-
Si3N4) is added as the bonding phase to form superhard nanocomposite
coatings. This “superhard” property has resulted in an enhanced cutting
performance. For instance, Jilek et al. (2004) found that the life of nc-
AlTiN/a-Si3N4 coated tools was greatly longer than that of AlTiN coated
tools in drilling a steel. More examples can be found in the review by
Veprek and Veprek-Heijman (2008). However, it should be noted that the
hardness of nanocomposite coated tools is significantly affected by the
grain size of nanocrystals (Musil, 2012).

Effect of improved coating structures


The initial development on coating structures has been focused on increas-
ing the number of layers to form multilayered coatings. Wang (2000) found
MACHINING SCIENCE AND TECHNOLOGY 483

that multiple layered coatings (TiC/Al2O3/TiN) reduced the cutting forces,


although the reduction was marginal under lighter cutting conditions.
Schintlmeister et al. (1981) also reported that TiC/TiN coated tools showed
smaller flank wear than monolayer-TiN coated tools.
Comparing to monolayered coatings, multilayered costing designs
allow the coatings to have an enhanced overall performance. First, an inner
layer with good bonding to the substrate (typically TiC) can reduce the
spalling of the coating. Second, intermediate layers are able to provide the
coating with a number of additional functions that are considered benefi-
cial to machining operations. For example, the multi-layered design of
TiC/TiCN/TiN has been able to reduce the gradient of thermal expansion
within the entire coating-substrate system, and successfully eliminated the
occurrence of thermal-induced cracks in the milling of a structural steel
(Schintlmeister et al., 1981).
A further improvement based on the conventional multilayered coatings
is the application of superlattice coatings. A typical superlattice structure is
shown in Figure 2, in which it can be seen that superlattice coatings are
nano-scale multilayers composed of two or more different alternating layers
with a superlattice period normally less than 10 nm (Hovsepian et al.,
2006). In contrast with conventional multilayered structures, superlattice
structures are able to provide the coatings with a super-high hardness rang-
ing from 40 to 60 GPa (Hovsepian et al., 2000). An explanation for the

Figure 2. Cross-sectional TEM micrograph of TiAlN/VN superlattice coatings (with 3.03 nm


period) (Hovsepian et al., 2006).
484 Y. CHEN ET AL.

greatly increased hardness is given by Yashar and Sproul (1999) and can be
summarized due to two factors: a limitation to dislocation mobility within
the superlattice layers and across layer interfaces caused by the image
forces, and the Hall-Petch strengthening effect.
The super-high hardness of the superlattice coatings has promoted
greatly the abrasion resistance of the cutting tools. Additionally, the super-
lattice structure may lead to a higher toughness and a better crack propaga-
tion resistance (Klocke and Krieg, 1999). As a result, superlattice hard
coatings have been very successful in reducing tool wear in machining
processes. Hovsepian et al. (2000) found that the life of TiAlYN/VN super-
lattice coated tools was more than 20 times longer than that of uncoated
cemented carbide tools and nearly three times longer than that of monolay-
ered TiCN coated tools in the milling of a steel.

Effect of soft coatings and relevant mechanisms


Due to environmental considerations, it has been widely advocated that the
use of pollutive cutting fluids should be eliminated. Since the tool-chip fric-
tion would be much more severe without the supplement of coolants and
lubricants, soft tool coatings have been used for “dry” machining opera-
tions. Unlike the hard coatings described above, soft coatings are based on
different philosophies on wear resistance. During the tool-chip contact in
machining, hard coatings usually aim at retaining minimum changes in
response to external loads, while soft coatings act as a solid lubricant with
small shear strength along the sliding direction to improve the surface
tribological performance.
Molybdenum disulfide (MoS2) is the most widely-used soft coating
material for cutting tools. This popularity attributes greatly to its special
microstructure, whereby MoS2 is formed by stacking “sandwiches” of a
layer of molybdenum (Mo) atoms arranged in a hexagonal array with two
layers of sulfur atoms (S) surrounded at equal distance. Between the
“sandwiches”, only Van der Waals forces exist, which results in easy shear-
ing along the basal plane of the hexagonal crystalline structures (Teer
et al., 1997).
However, the initial applications of MoS2 to cutting tool costings
have not achieved what was expected. Liu et al. (1999) found that MoS2
coating was able to extend tool life in the cutting of steels only when the
cutting parameters were selected in certain ranges, and the beneficial effect
was not significant. This unsuccessful experience with MoS2 coatings was
probably due to the property of MoS2 that degrades when getting contact
with humid air, leading to an increased coefficient of friction (Renevier
et al., 2001).
MACHINING SCIENCE AND TECHNOLOGY 485

To improve the effectiveness of MoS2 coatings in humid environments,


some metal elements, e.g. Ti (Renevier et al., 2001) and Zr (Deng et al.,
2008), are incorporated into the coating, forming MoS2/metal composite
coatings. The mix of metal elements has also enhanced the density and
hardness of the coating, while still retaining a low coefficient of friction
(Teer et al., 1997). As a result, the cutting performance of MoS2/metal
coated tools has been markedly improved when compared to the original
MoS2 costings, as demonstrated by the work of Renevier et al. (2000) in
the milling of an AISI 304 steel by MoS2/Ti coated tools.
Although MoS2/metal soft coatings have achieved good machining per-
formance in some cases, it seems that these coatings are only effective in
low-speed cutting processes while in high-speed machining, its effectiveness
was only marginal. This has been reported by Deng et al. (2008), in which
they found that for cutting speeds less than 120 m/min, the flank wear and
cutting forces of MoS2/Zr coated tools were significantly lower than those
of uncoated tools, but when the cutting speed was above 120 m/min, both
the flank wear and cutting forces increased sharply (as shown in Figure 3).
The authors further reported that a number of cracks were found on the
coatings at high temperatures during high-speed cutting, and these cracks
would initiate the loss and spalling of the coating. Another possible cause
for the undesirable performance was reported by Song et al. (2017) who
believed that MoS2 would be oxidized at the temperature of around 400  C
and transform to MoO3, leading to high friction at the tool-work and tool-
chip interfaces.

Figure 3. Flank wear of MoS2/Zr coated tools and uncoated tools versus cutting speed in steel
machining (Deng et al., 2008).
486 Y. CHEN ET AL.

To further improve the performance of soft coatings, especially for high-


speed machining, tungsten (W) has been added to the coating materials as
a replacement of Mo, and such coating materials are known as WS2-based
composites. Owing to the good thermal stability and oxidation resistance
provided by W, WS2-based soft coatings have performed successfully in
reducing the friction and wear in machining, even in high-speed cutting
operations. Relevant examples can be found in the study by Li et al. (2015)
and Lian et al. (2013).

Cutting tool surface modification by high energy beams


Surface modifications have been made by three high energy sources, i.e. laser
beams, ion beams and electron beams. Among them, ion implantation and
high current electron beam (HCEB) treatment have been successfully applied
to surface modification to increase the wear resistance (Dearnaley et al.,
1985; Conrad et al., 1987). The improvements in the cutting performance
through cutting tool surface modifications by ion implantation and high cur-
rent pulsed electron beam (HCPEB) treatment are discussed below.

Effect of ion implantation and relevant mechanisms


Wear reduction of cutting tools by ion implantation is mainly caused by
the improved mechanical properties within the surface and subsurface
region of the cutting tools. However, to obtain such improvements is not
an easy task since the modification depends largely on the target material
composition and microstructure and its interaction with the implanted ele-
ments. Nevertheless, for the most popularly used cutting tool material,
cemented tungsten carbide, evidence has shown that both the hardness and
tribological properties were enhanced by ion implantation.
Dearnaley et al. (1985) found from the nitrogen implantation process
that at an implantation temperature above 200  C, a significant increase in
the surface hardness on WC-Co was achieved; however, at a relatively low
processing temperature (less than 200  C) nitrogen implantation would
reduce the WC-Co surface hardness. Sun et al. (1997) found from a friction
test by sliding contact that a longer low-friction sliding distance and less
adhesion were achieved with nitrogen implanted WC-Co samples. They
further showed that nitrogen implantation could give rise to the formation
of a carbon-rich layer within the subsurface of the WC-Co target, and such
a layer was responsible for the reduced friction.
The MEVVA technique has been used to modify cutting tool surfaces.
Surprisingly, the implantation of some specific metal elements has led to
more significant promotion on the cutting tool surface hardness than
MACHINING SCIENCE AND TECHNOLOGY 487

nitrogen, as reported by Fu et al. (2005), in which the researchers


have achieved up to a 250% increase on the hardness by Mo þ W co-
implantation, compared to 150% by nitrogen. However, the reduction of
the frictional coefficient by Mo þ W co-implantation on WC-Co-based cut-
ting tool materials was about 19%, compared to 25% by nitrogen implant-
ation (Fu et al., 2005).
Although ion implantation can improve the mechanical properties of the
surface of cutting tools, the machining performance of the ion implanted
cutting tools has not been as good as expected. It was found that both
nitrogen (Saklakoglu et al., 2007) and metal implantations (Treglio et al.,
1993) were only effective in low-speed cutting processes. This is probably
because ion implantation is not able to comprehensively increase the mech-
anical properties of the cutting tool surfaces at high machining tempera-
tures. During high-speed cutting, thermal-induced wear mechanisms, such
as adhesive wear, diffusion wear, and oxidation, can be dominating, but no
evidence has shown that ion implantation can lead to an increase in the
high-temperature wear resistance of cutting tools.
Some exciting outcome has been achieved by a combined application of
surface coating and ion implantation. In such attempts, ion implantation
was either a pretreatment process to form a functional interfacial layer
between the substrate and the outer coating (Shum et al., 2005), or a post-
treatment process to enhance the mechanical properties (such as hardness
and elastic modulus) and tribological properties of the coating (Manory
et al., 1998; Deng et al., 2012). A good example of such a combined appli-
cation has been reported by Shum et al. (2005) for high-speed machining.
The researchers found that using titanium implantation as a pretreatment
on WC-Co substrate could result in a greatly prolonged tool life in the
high-speed turning of an M238 mound steel, as compared to the process
using cutting tools with only a TiAlN coating.

Effect of HCPEB treatment and relevant mechanisms


Similar to ion implantation, HCPEB treatment is able to change the mech-
anical and metallurgical properties of the treated surface through a number
of fast heating-quenching cycles. These changes are beneficial to the
machining applications have mainly been reported on cemented carbides
(WC-Co-based), and through an increased cutting tool surface hardness
and reduced friction coefficient (Ivanov et al., 2000; Uglov et al., 2012; Hao
et al., 2013; Xu et al., 2013b). Ivanov et al. (2000) reported that the
major cause for the increased hardness in HCPEB treatment was by the
solid-solution hardening effect within the Co binder phase. However, this
hardening effect appears to be very limited as compared to that by cutting
488 Y. CHEN ET AL.

tool surface coating and ion implantation. Based on the work of Xu et al.
(2013b), the maximum surface microhardness of WC-Co targets reached
21.8 GPa after HCPEB treatment, which was 36% higher than the
untreated material.
By contrast, the improvement in the tribological properties of WC-Co-
based carbides by HCPEB treatment is significant. In the work by
Uglov et al. (2012), it was found the friction coefficients of WC-Co-based
samples were decreased by 2 to 3.5 times after HCPEB irradiation. These
low friction coefficients were believed to be a result of a self-lubrication
effect provided by a graphite layer formed during the HCPEB surface treat-
ment. Hao et al. (2013) have detected the formation of such a graphite
layer consisting of nano graphite particles through surface analysis.
There have been reported studies of the cutting performance using
the HCPEB treated cutting tools. Interestingly, HCPEB irradiation has
been found to be particularly effective in reducing the tool wear in high-
speed machining, as in the work by Ivanov et al. (2000) for the turning of
a 40X steel and by Proskurovsky et al. (1998) for the cutting of a 40 steel.
However, the cutting performance of HCPEB treated tools in relatively
low-speed cutting operations is not encouraging. As reported by Ivanov
et al. (2000), no obvious difference was found on the tool wear rate
between the treated and untreated tools when the cutting speed was lower
than 200 m/min in steel machining.
It should be noted that as a thermal-based surface modification tech-
nique, HCPEB irradiation can cause thermal-induced damages on the
target surface. A typical example of such surface damages is micro-cracks.
A large number of micro-cracks can be generated on the target surface by
cyclic heating and quenching through high beam intensity and/or a large
number of pulses (Ivanov et al., 2000; Xu et al., 2013b). Typical micro-
cracks generated due to excessive pulses are shown in Figure 4. The forma-
tion of such micro-cracks can significantly affect the cutting performance
of the cutting tools. Therefore, the application of electron beams in HCPEB
surface treatment has to be properly controlled to achieve the desired sur-
face properties.

Cutting tool surface texturing


Micro-scale surface textures in various shapes have been able to be fabri-
cated through a wide range of techniques including laser micro-machining
(Wan et al., 2011; Manickam et al., 2013), electric discharge machining
(EDM) (Koshy and Tovey, 2011; Kim et al., 2016), photolithography
(Obikawa et al., 2011), and micro-grinding (Xie et al., 2013), while nano-
scale textures are basically created by femtosecond pulsed lasers in practical
MACHINING SCIENCE AND TECHNOLOGY 489

Figure 4. Micro-cracks generated on the HCPEB irradiated WC-Co surface after (a) 1, (b) 5, (c)
20, (d) 35 pulses (Xu et al., 2013b).

applications (Li and Wang, 2017). The mechanisms of surface texturing in


reducing friction and wear as well as other cutting performance quantities
in machining are discussed below.

Mechanisms of friction and wear reduction by surface texturing


in machining
Micro-pool lubrication
The contact between the tool and chip in machining is distinctly different
from the conventional solid-solid sliding contacts. This is due to not only
the extremely high loads acting on the tool rake face, particularly at the
area near the main cutting edge, but also the nonuniform distribution of
such loading, high sliding speeds, and high temperatures at the contact.
The real area of contact near the main cutting edge can approach 100% of
the nominal contact area, indicating a nearly complete asperity contact,
and neither gaseous nor liquid lubricant could penetrate into such contact
area (Childs, 2000). At the rear of tool-chip contact, limited volume of
lubricant may infiltrate along the surface channels which are probably
formed by abrasion through capillary actions, and react with the solid to
form a low-friction layer (Childs, 2000). Consequently, even for wet cutting
490 Y. CHEN ET AL.

Figure 5. A schematic of hydrodynamic effect (Wang et al., 2003).

operations, the lubrication regime for conventional cutting tools on the


tool rake face may still be a combination of dry and boundary lubrication.
Surface texturing is able to transit the lubrication regime to a mixed lubri-
cation, since surface textures may either help to imbibe lubricant by provid-
ing the tool rake face with a superwet behavior (discussed in the next
section) or act as a reservoir to retain lubricant. Within the mixed lubrica-
tion regime, compressive loads and relative sliding actions can squeeze the
trapped lubricant out from the surface pockets by a joint hydrodynamic and
hydrostatic effect. The escaped lubricant can then form a thin interfacial film
between the two sliding surfaces and, hence, reduce the friction and wear.
This process is known as micro-pool lubrication (Lo and Wilson, 1999).
A schematic of the hydrodynamic effect is shown in Figure 5 (Wang
et al., 2003). When the upper surface slides with a velocity of U with
respect to the lower surface, the trapped lubricant also moves in the same
direction. The generation of the hydrodynamic effect is mainly due to the
increase of the fluid pressure at the converging portion. Although the pres-
sure decreases at the divergence portion on the other side of the symmet-
rical slope, it cannot reach an ideal anti-symmetrical value because of
the cavitation effect (Wang et al., 2003). Hence, the lowest pressure at this
portion is the cavitation pressure. As a result, considerable hydrodynamic
pressure can be generated, especially when the sliding speed is high, caus-
ing continuous permeation of lubricant into the contact interface.
The generation of hydrostatic pressure results from the compression
between two contacting surfaces. Under a certain compressive load, the
high spots on both surfaces can be depressed and flattened (Kasuga and
Yamaguchi, 1968). This forces the volume of surface pockets to be reduced,
causing an increase in the liquid (lubricant) hydrostatic pressure. When
this hydrostatic pressure exceeds the interfacial pressure on the surround-
ing plateau, the lubricant is squeezed into the contact interface.
MACHINING SCIENCE AND TECHNOLOGY 491

Figure 6. A schematic of liquid contact angle (notated as h).

In addition, Lei et al. (2009) have pointed out that there is usually a large
mismatch in the thermal expansion rates of lubricant and tool material.
Such mismatch may contribute to the increase of fluid pressure at high
temperatures, which may be another factor causing the escape of trapped
lubricant from the micro-pools.

Improvement of surface wettability for lubricant


The wettability refers to the ability of a solid surface to spread or agglomer-
ate a liquid on it. The degree of wettability is quantified by a liquid contact
angle, which is defined as the angle subtended at the triple-phase contact
line formed by a fluid-air-solid interface, as shown in Figure 6 (notated as
h). An extreme state is that the contact angle is very small (typically less
than 5 ), and the liquid may propagate spontaneously on the solid surface
to cause wet on it. Such a state is usually known as superlyophilicity and
the surface is called superwetting surface.
A good surface wettability of cutting tools is undoubtedly beneficial to
the cutting performance, as cutting fluid can be imbibed into the tool-chip
contact area more easily to provide an enhanced lubrication effect, a
reduced tool-chip friction, and a smaller tool wear rate. As wettability is
significantly influenced by not only the environmental factors and the
liquid properties, but also the topography of the solid surface, surface tex-
turing is able to improve surface wettability (Bico et al., 2001; Cunha et al.,
2013). A well-known model developed by Wenzel (1936) indicates that the
cosine of the apparent contact angle is proportional to the value of surface
roughness (defined as the ratio of the actual solid surface area and the pro-
jected surface area). However, Wenzel’s model is only able to illustrate the
static state of the bulk droplet, and a more important factor involved in
superwet-lubrication, i.e. the spontaneous spreading action of liquid within
the superlyophilic regime (including the criterion and the spreading rate),
was not given.
Bico et al. (2002) developed another model, in which a particular
interest was focused on the superwetting behavior. As analyzed by the
authors, the spreading of the liquid front is dominated by the evolution of
the solid-liquid-vapor interfacial free energies and the viscous dissipation
492 Y. CHEN ET AL.

effect. Properly designed surface textures can keep the variation of system
interfacial energy to be negative while the liquid front propagates, and
it results in a spontaneous permeation within the forest of micro/nano sur-
face structures or along the groove-shaped micro-channels. Further, both
the propagation rate and the transmission flux are greatly dependent on the
texture design. This theory has been supported by experimental works
(Vorobyev and Guo, 2009; Vorobyev and Guo, 2010). Vorobyev and Guo
(2010) found through laser surface texturing on a silicon sample that micro-
scale grooves covered with numerous nano-scale protrusions and cavities
could be formed on the target surface, and water was able to permeate on
the textured surface at a considerable rate even towards the uphill direction,
indicating a highly superhydrophilic performance. A similar phenomenon
has also been found on metal surfaces (Vorobyev and Guo, 2009).
Furthermore, it has been confirmed that surface textures would indeed
facilitate the penetration of liquid lubricant in practical cutting operations,
although a quantitative description of the wettability of the textured surface
was not given (Koshy and Tovey, 2011). In this work, a comparison was
made between textured tools and nontextured tools in the turning of a
steel under an intermittent cutting fluid supplying condition, as shown in
Figure 7. For the nontextured tools, the tool-chip friction with and without
cutting fluid supply shows a similar fashion, indicating a very limited lubri-
cation effect. By contrast, for the textured tools, the tool-chip friction
decreases significantly when cutting fluid was supplied.

Trapping of abrasive particles


During sliding contact, abrasive wear is a very common wear mechanism
and has been divided into two groups, i.e. two-body abrasion and three-
body abrasion. The former refers to the abrasion by constrained abrasives

Figure 7. Friction angle of: (a) textured tools; and (b) nontextured tools (shaded regions denote
the time duration with cutting fluid supply) (Koshy and Tovey, 2011).
MACHINING SCIENCE AND TECHNOLOGY 493

with only sliding action, while the latter is where abrasives are free to roll.
Debris generated from previous abrasive wear or the breaking of interfacial
adhesive bonds (pulled-out particles) would be the main cause of three-
body abrasion wear. The surface pockets formed by the fabricated textures
can trap debris in both dry and lubricated friction conditions in machining
(Sugihara and Enomoto, 2013). The entrapment of hard particles can pre-
vent them from further sliding in the contact interface.

Reduction of tool-chip contact area


In a number of studies (Wu et al., 2012; Deng et al., 2013; Zhang et al.,
2015) it has been proposed that surface textures could reduce the tool-chip
friction by decreasing the tool-chip contact area, based on the assumption
that a constant average shear stress applied on the tool rake face. However,
the reported studies have essentially assumed that both the tool surface and
the chip surface were perfectly smooth (except for the textured area), which
may not be the case in practice. More and more evidence has shown that
the behavior of solid-solid contact and friction should be considered as a
micro- and/or nano-scale problem, so that the micro- and nano-scale sur-
face roughness may play a role in affecting the contact and friction behav-
iors of two surfaces (Greenwood and Williamson, 1966; Black et al., 1993).
As such, if and how cutting tool surface textures can reduce the tool-chip
contact area requires further investigations.
Furthermore, surface textures may indeed reduce the nominal tool-chip
contact area, but a reduced nominal contact area may cause an increase in
the compressive stress at the tool-chip interface, which in turn increases
the density of contact points or the deformation of the asperities
(Greenwood and Williamson, 1966; Uehara et al., 2004). As a result, the
reduced nominal contact area considered at a macro-scale may be traded
off by the increased asperity contact area at micro- and/or nano-scales, so
that the contact area actually reduced may be much smaller than intuitively
believed, except for the contact near the main cutting edge where such a
reduction may be negligible, as the actual contact area can be intrinsically
close to the nominal one due to the extremely high compressive load.
Unfortunately, considering the entire tool-chip contact, no reported studies
were found which clearly express the relationship between the textured
area and the reduced tool-chip contact area.

Effect of surface textures on cutting performance


Due to the effectiveness of surface textures in reducing friction and wear, con-
siderable studies have been reported to apply this technique to cutting tools.
Some representative studies reported recently are summarized in Table 1. The
494

Table 1. Summary of recent and representative studies on the cutting performance of surface textured cutting tools.
Reference Process Tool/workpiece material Texture characteristics Cutting conditions Major factors concerned
(Kawasegi et al., 2017) Cutting  Tool material: Single crys-  Dimension: micro-scale  Cutting speed: 130 m/min  Groove orientation
tal diamond  Shape: groove  Feed rate: 10 lm/stoke  Groove width
 Workpiece material: A5052  Orientation: parallel/perpen-  Depth of cut: 3 lm  Depth of cut
Y. CHEN ET AL.

aluminum alloy, NiP dicular to the cutting edge  Lubrication: dry/wet  Lubrication condition
 Placement: tool rake face
(Vasumathy and Turning  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 150 m/min  Groove orientation
Meena, 2017) cemented carbide  Shape: groove  Feed rate: 0.16 mm/rev
 Workpiece material: AISI  Orientation: parallel/perpen-  Depth of cut: 1 mm
316 austenitic stain- dicular to the main cutting  Lubrication: dry
less steel edge, and diagonal
 Placement: tool rake face
(Gajrani et al., 2018) Turning  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 55–125 m/min  Groove orientation
cemented carbide  Shape: groove and  Feed rate: 0.04–0.28 mm/rev  Texture shape
 Workpiece material: AISI H- round dimple  Lubrication: dry/solid lubricant  Solid lubricant
13 steel  Orientation: parallel/perpen-  Feed rate
dicular to the main cut-  Cutting speed
ting edge
 Placement: tool rake face
(Sugihara and Milling  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 200 m/min  Texture shape
Enomoto, 2017) cemented carbide  Shape: groove and  Feed rate: 0.2 mm/tooth  Texture area ratio
 Workpiece material: round dimple  Depth of cut: 2 mm  Lubrication condition
Medium carbon steel  Orientation: parallel to the  Lubrication: wet/dry/paste
main cutting edge
 Placement: tool rake face
(Durairaj et al., 2018) Tube turning  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 300 m/min  Dimple diameter
cemented carbide  Shape: round dimple  Depth of cut: 50 lm  Dimple depth
 Workpiece material:  Placement: tool rake face  Cutting length: 7.5 m  Texture pitch
Al6061 alloy  Lubrication: wet  Distance between textures
and the main cutting edge
(continued)
Table 1. Continued.
Reference Process Tool/workpiece material Texture characteristics Cutting conditions Major factors concerned
(Sasi et al., 2017) Tube turning  Tool material: T-42 high  Dimension: micro-scale  Cutting speed: 20–30 m/min  Cutting speed
speed steel  Shape: round dimple  Depth of cut: 0.175 mm  Solid lubricant
 Workpiece material: Al7075-  Placement: tool rake face  Cutting width: 1.25 mm
T6 alloy  Lubrication: dry/solid lubricant
(Sugihara et al., 2017) Tube turning  Tool material: CBN  Dimension: micro-scale  Cutting speed: 300 m/min  Groove orientation
 Workpiece material: Inconel  Shape: groove  Feed rate: 0.1 mm/rev  Distance between textures
718 alloy  Orientation: parallel/perpen-  Lubrication: wet and the main cutting edge
dicular to the main cut-  Flank face textures
ting edge
 Placement: tool flank face
(Liu et al., 2017) Turning  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 120 m/min  Groove orientation
cemented carbide  Shape: groove  Feed rate: 0.051 mm/rev  Flank face textures
 Workpiece material: Al2O3  Orientation: parallel/perpen-  Depth of cut: 0.4–0.8 mm
green ceramic dicular to the main cutting  Lubrication: dry
edge, and diagonal
 Placement: tool flank face
(Liu et al., 2018) Turning  Tool material: WC-Co  Dimension: micro-scale  Cutting speed: 120 m/min  Groove width
cemented carbide  Shape: groove  Feed rate: 0.051 mm/rev  Groove spacing
 Workpiece material: Al2O3  Orientation: parallel to the  Depth of cut: 0.8 mm  Distance between textures
green ceramic main cutting edge  Lubrication: dry and the main cutting edge
 Placement: tool flank face  Flank face textures
(Zhang et al., 2017) Turning  Tool material: WC-Co  Dimension: nano-scale  Cutting speed: 40–200 m/min  Cutting speed
cemented carbide  Shape: periodical  Feed rate: 0.1 mm/rev  Nano-textures
 Workpiece material: nano-ripples  Depth of cut: 0.3 mm  Hard coating
Hardened steel  Placement: tool rake face  Lubrication: dry/solid lubricant  Solid lubricant
(Hao et al., 2018) Turning  Tool material:  Dimension: micro-scale  Cutting speed: 30–120 m/min  Cutting tool surface
Polycrystalline diamond  Shape: groove  Feed rate: 0.16 mm/rev wettability
 Workpiece material: Ti-6Al-  Orientation: parallel and  Depth of cut: 0.2 mm  Cutting speed
4V alloy perpendicular to the main  Lubrication: wet
cutting edge
MACHINING SCIENCE AND TECHNOLOGY

 Placement: tool rake face


495
496 Y. CHEN ET AL.

effect of surface textures on the cutting performance is dependent on the tex-


ture geometrical parameters (such as dimension, shape, and orientation) in
addition to cutting conditions, and the reported studies have enabled to look
into the effect of individual factors. These are reviewed below.

Effect of texture dimension


Nano-scale textures. Nano-scale textures (nano-textures) are mainly referred
to the laser induced periodical surface structures (LIPSS), with a dimension
of several hundred nanometers, as shown in Figure 8. Although in some
published works (Kawasegi et al., 2009) nano-textures are formed simultan-
eously with, typically, micro-grooves, those surface textures are regarded as
multi-scale micro/nano textures and will be discussed later in the paper.
It has been found that when machining aluminum alloys, nano-textures
were unable to reduce the adhesion between the tool and work material (at

Figure 8. Nano-scale textures fabricated using a femtosecond laser (Sugihara and


Enomoto, 2012).
MACHINING SCIENCE AND TECHNOLOGY 497

the tool-work and tool-chip interfaces) in either wet (or lubricated)


(Sugihara and Enomoto, 2012) or dry (Sugihara and Enomoto, 2009) cut-
ting conditions, although they have shown some effect when applied
on coated cutting tools (Sugihara and Enomoto, 2012). This is probably
due to the fact that nano-textures are too small, and the volume of the
spontaneously propagated lubricant through the nano-channels may be
too limited to provide an effective pooled lubrication. Furthermore, due
to the high tendency of adhesion in the cutting of aluminum alloys, nano-
textures may be easily buried by the adhered work material and thus lose
their functions.
In the machining of hard materials such as steels, the performance of
nano-textured cutting tools has also been unsatisfactory. Enomoto et al.
(2012) found that nano-textured tools suffered from almost the same
degree of crater wear as those on nontextured tools in the milling of a steel.
This is probably because nano-textures are easily worn off as the cut-
ting progresses.
Combining surface modification with nano-textures has been reported to
give a good improvement in cutting performance (Deng et al., 2013). In this
work, a soft coating was made on nano-textured cutting tool surface which
reduced cutting forces and cutting temperatures. Combining both a TiAlN
hard coating and a soft WS2 soft coating with nano-textures has also been
successful in improving the cutting performance and the machined surface
quality in the turning of a hardened steel (Zhang et al., 2017).

Micro-scale textures. Micro-scale textures (micro-textures) are usually the


surface structures with a dimension ranging from several microns to several
hundred microns, as shown in Figure 9. Compared to conventional tools
and nano-textured tools, micro-textures were able to significantly reduce
adhesion between the tool and work in both wet and dry cutting of alumi-
num alloys (Kawasegi et al., 2009; Sugihara and Enomoto, 2012). As
claimed by Sugihara and Enomoto (2012), in the machining of aluminum
alloys using an uncoated tool, whether it is under a lubricated or dry con-
dition, the genesis phase of chip adhesion was difficult to prevent, but

Figure 9. Micro-scale textures: (a) dimple texture, (b) groove texture perpendicular to the main
cutting edge, (c) groove texture parallel to the main cutting edge (K€ummel et al., 2015).
498 Y. CHEN ET AL.

micro-textures might be large enough to reduce the growing phase of chip


adhesion by dividing the chip adhesion layer into separate parts and pre-
venting it from growing to an originally-tended areal nervation shape.
Through micro-texturing, a reduced wear rate has been observed in the
machining of steels (Lei et al., 2009; Kim et al., 2016), titanium alloys (Ling
et al., 2013; Xie et al., 2013), and a Ni-P alloy (Kawasegi et al., 2017). The
improved cutting performance by micro-textures may result from an
increased concave space for both retaining lubricant and trapping abrasive
particles, although the convex area which can contribute to bearing the
compressive stress and resisting mechanical wear may be decreased.

Multi-scale micro/nano textures. A promising result in improving the cutting


performance has been achieved by using multi-scale micro/nano textures
on the cutting tools, typically nano-structural arrays are imbedded on
micro-structures. By fabricating micro-wavy structures combined with
nano-textures on the micro-waves on WC-Co cemented carbide cutting
tools for the turning of an aluminum, Kawasegi et al. (2009) have shown
that micro/nano-textured tools yielded smaller cutting forces and friction
than micro-textured cutting tools. Further, the multi-scale textures were
found to be more effective in reducing chip adhesion than micro-textures.
It is believed that the reduced cutting force may be attributed to the
reduced work adhesion on the tool rake face.

Effect of texture shape


While various shaped surface textures have been attempted for cutting
tools, the most commonly used ones have been round dimples and grooves,
as shown in Figure 10, primarily because of their ease of fabrication.

Round dimples. A representative work to improve the cutting performance


of the tungsten carbide tools with round dimple textures for cold rolled
1045 steel is reported by Lei et al. (2009). For the textured tools, cutting

Figure 10. Rake face of a rounded-textured tool (a) and a grooved-textured tool (b) after a cut-
ting operation (Sugihara and Enomoto, 2017).
MACHINING SCIENCE AND TECHNOLOGY 499

fluid was prefilled (oil) into the surface pockets, while for the conventional
tools, flood cooling was applied. It was found that all the cutting force
components, as well as the tool-chip friction, were reduced by the surface
modification. Apparently, this is attributed to the effective micro-pool
lubrication and a reduction in the tool-chip contact area. In addition, the
round dimples filled with cutting fluid has caused an early separation of
the chip from the tool rake face, leading to a much shorter tool-chip con-
tact length (reduced by 28%). According to the study by Zhao et al. (2007),
round dimple textures may impede the spreading of liquid on a solid
surface, so that with this form of textures, it may be difficult to realize a
spontaneous penetration of cutting fluid in machining.

Grooves. A notable difference between grooved and rounded patterns is that


grooves are connected in one direction, while round dimples are completely
disconnected. Grooves may cause easier penetrations of cutting fluid in wet
cutting operations to reduce tool-chip and tool-work friction, as confirmed
by the experimental work of Obikawa et al. (2011) and Sugihara and
Enomoto (2017), although the advantage of grooved over rounded textures
was not found to be significant.
It should be realized that the edge of surface textures may act as a
micro-cutter, and associated micro-cutting actions may take place between
the texture edges and chips during chip flow (Sugihara and Enomoto,
2017). Such micro-cutting actions may lead to an increased tool-chip fric-
tion. Figure 10 shows the rake face of a rounded-textured tool and a
grooved tool after a machining operation (Sugihara and Enomoto, 2017). It
is shown that more severe micro-cutting took place on the groove-shaped
textures, and the “chips” generated in such micro-cutting could bury the
grooves, which in turn could suppress the penetration of cutting fluid or
increase the actual tool-chip contact area. This may be a reason why the
advantage of grooved textures over rounded ones was very marginal
(Obikawa et al., 2011; Sugihara and Enomoto, 2017).
A comprehensive investigation into the micro-cutting process for micro-
grooved cutting tools has been reported by Duan et al. (2018), where an
analytical approach was used to predict the formation of chips on the
grooved surface of the cutting tools. It was found that either increasing the
edge radius of the grooves or applying a higher cutting speed could effect-
ively reduce the formation of the chips. Nevertheless, since the dimension
of the grooves is at the micro-meter scale, increasing the groove edge
radius may require a much more complex micro-machining process for
the groves, so that enhancing the cutting speed is more practical, although
it may induce some other issues, such as the impedance of cutting fluid
penetration (Cassin and Boothroy, 1965). Further investigations are yet
500 Y. CHEN ET AL.

required for minimizing the formation of chips from the micro-cutting by


the groove edges on the cutting tool.

Effect of texture orientation


Texture orientation is a parameter associated with grooved textures. In this
work, the direction of textures is defined with respect to the chip flow dir-
ection. The reported studies are essentially focused on parallel and perpen-
dicular grooves, where the latter were found to always outperform the
former (Kawasegi et al., 2009; Sugihara and Enomoto, 2009). Kawasegi
et al. (2009) found that work material is more likely to be pressed into the
grooves if the chip flow direction is consistent with the groove orientation.
Such an action would cause either an increased tool-chip contact area or
the burying of surface textures. Moreover, Kawasegi et al. (2017) found
that the press-in of the chip might be more severe when cutting a high-
strength material, since higher cutting forces may be required which would
lead to a greater compression. For multi-scale micro/nano textures, the
orientation of nano-textures was found to have no significant influence on
cutting performance (Kawasegi et al., 2009).
It is worth mentioning that grooved textures have an anisotropic wetting
property (Vorobyev and Guo, 2010) and promote the permeation of liquid
in the direction parallel to the grooves. Because the penetration of cutting
fluid is usually directional (Godlevski et al., 1997), connecting the source of
cutting fluid to the tool-chip contact area using micro-grooves can provide
a good effect in imbibing the cutting fluid.

Effect of cutting speed


The effect of cutting speed on the effectiveness of cutting tool surface tex-
tures in improving the cutting performance can be discussed in two
aspects. First, a higher cutting speed can encourage the formation of
micro-pool lubrication. This has been demonstrated by Bech et al. (1999)
through experiments, and further by Shimizu et al. (2001) through theoret-
ical analysis. An enhanced micro-pool lubrication effect by an increased
cutting speed can improve the availability of cutting fluid in the tool-chip
contact interface and reduce the friction force. Second, an increased cutting
speed can impede the penetration of cutting fluid into the tool-chip contact
area. This has been confirmed for both conventional cutting tools (Cassin
and Boothroy, 1965) and textured cutting tools (Koshy and Tovey, 2011).
Koshy and Tovey (2011) further show that the high tool-chip interface
temperatures at high cutting speeds can cause lubricant to breakdown.
Owing to the two opposite effects mentioned above by cutting speed, it
is not unreasonable to conclude that there should be a range of cutting
MACHINING SCIENCE AND TECHNOLOGY 501

speeds within, with which the effectiveness of cutting tool surface textures
in improving the cutting performance is promoted. This is consistent with
the finding by Obikawa et al. (2011) who arrived at the conclusion after
comparing their experimental data with those of Kawasegi et al. (2009).

Concluding remarks
The development toward high speed (or high efficiency) and high precision
machining, and the need to machine increasingly difficult-to-machine
materials, requires constant improvements in the performance of cutting
tools. While new cutting tool materials and high-performance cutting tool
designs are being developed along with the optimization of cutting condi-
tions and cutting fluid applications, modifications to the mechanical prop-
erties and topography of cutting tool surfaces have been found to be an
effective way to improve the cutting performance, such as to reduce friction
and tool wear. The commonly used approaches to modify the cutting tools
surfaces include surface coating, high energy beam treatment, and surface
texturing. A review of how these surface modification techniques improve
the various cutting performance quantities as well as their advantages and
limitations has been presented.
While cutting tool surface coating or heat treatment may be considered
as a passive way to improve the performance of cutting tools, surface tex-
turing is able to improve the lubrication conditions and trap hard debris to
reduce three-body abrasive wear in machining. It may be noted that surface
treatment by high energy beams is usually not as effective as surface coat-
ing in improving the cutting tool performance. This is due to the fact that
coating materials and structures can be relatively easily optimized according
to the tool material composition and the machining process needs, while
the high energy beam heat treatment depends heavily on the interaction
between the target material and the energy beam. However, both surface
coating, particularly by CVD, and HCPEB treatment involve high tempera-
tures. If not controlled properly, these surface modification processes can
cause undesirable material structural changes, such as the generation of a
brittle phase known as the g-phase to WC-Co-based cemented carbide,
and thermal induced micro-cracks.
By contrast, surface texturing is not aimed to change the mechanical,
chemical or metallurgical properties of the tool material, but to reduce
tool-chip and tool-work friction and tool wear based on different mecha-
nisms as presented in this article. Although the cutting performance of
purely surface-textured cutting tools has often been shown to be not as
good as that of surface coated tools, further studies are required to opti-
mize the texture design for cutting tools and understand the friction and
502 Y. CHEN ET AL.

wear mechanisms of surface textured tools, so that the full potential of this
surface modification technique in improving the cutting performance can
be explored. Moreover, it has been found that the combined use of surface
texturing and surface coating (including both hard coatings (Mishra et al.,
2018; Zhang et al., 2018) and soft coatings (Lian et al., 2018; Xing et al.,
2018)) is a feasible avenue to further improve the cutting performance.
However, the cost associated with such a combined application needs to be
justifiable with the improved cutting performance.
It should be realized that the possible surface and/or sub-surface dam-
ages, and the possible reduction of tool edge strength by surface texturing,
should be minimized by selecting an appropriate texturing technique and
optimizing the texture dimension, orientation, shape and pattern. For
instance, conventional laser ablation may cause thermal induced micro-
cracks and residual stresses which undoubtedly reduce the reliability and
performance of the cutting tools. Ultra-short pulsed laser texturing is a
technique to realize damage-free or near damage-free “cold” material abla-
tion at micro- and nano-scales. Research on applying this technique to sur-
face modification for cutting tools and other engineering surfaces is a
promising direction.

ORCID
Jun Wang http://orcid.org/0000-0003-1699-3687

References
Aralan, A.; Masjuki, H. H.; Kalam, M. A.; Varman, M.; Mufti, R. A.; Mosarof, M. H.;
Khuong, L. S.; Quazi, M. M. (2016) Surface texture manufacturing techniques and tribo-
logical effect of surface texturing on cutting tool performance: A review. Critical Reviews
in Solid State and Materials Sciences, 41(6): 447–481.
Bech, J.; Bay, N.; Eriksen, M. (1999) Entrapment and escape of liquid lubricant in metal
forming. Wear, 232(2): 134–139.
Bico, J.; Thiele, U.; Quere, D. (2002) Wetting of textured surfaces. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 206(1–3): 41–46.
Bico, J.; Tordeux, C.; Quere, D. (2001) Rough wetting. Europhysics Letters, 55(2): 214–220.
Black, A.; Kopalinsky, E.; Oxley, P. (1993) Asperity deformation models for explaining the
mechanisms involved in metallic sliding friction and wear—a review. Proceedings of the
Institution of Mechanical Engineers, Part C: Journal of Mechanical Engineering Science,
207(5): 335–353.
Bouzakis, K.-D.; Michailidis, N.; Skordaris, G.; Bouzakis, E.; Biermann, D.; M’Saoubi, R.
(2012) Cutting with coated tools: Coating technologies, characterization methods and
performance optimization. CIRP Annals-Manufacturing Technology, 61(2): 703–723.
Cassin, C.; Boothroy, G. (1965) Lubricating action of cutting fluids. Journal of Mechanical
Engineering Science, 7(1): 67–81.
MACHINING SCIENCE AND TECHNOLOGY 503

Chen, M.; Jian, X.; Sun, F.; Hu, B.; Liu, X. (2002) Development of diamond-coated drills
and their cutting performance. Journal of Materials Processing Technology, 129(1–3):
81–85.
Childs, T. (2000) Metal machining: Theory and applications, Arnold, London, UK.
Chowdhury, S.; de Barra, E.; Laugier, M. (2004) Study of mechanical properties of CVD
diamond on SiC substrates. Diamond and related materials, 13(9): 1625–1631.
Conrad, J. R.; Radtke, J.; Dodd, R.; Worzala, F. J.; Tran, N. C. (1987) Plasma source ion-
implantation technique for surface modification of materials. Journal of Applied Physics,
62(11): 4591–4596.
Cunha, A.; Serro, A. P.; Oliveira, V.; Almeida, A.; Vilar, R.; Durrieu, M.-C. (2013) Wetting
behaviour of femtosecond laser textured Ti–6Al–4V surfaces. Applied Surface Science,
265: 688–696.
Dearnaley, G.; Minter, F.; Rol, P.; Saint, A.; Thompson, V. (1985) Microhardness and nitro-
gen profiles in ion implanted tungsten carbide and steels. Nuclear Instruments and
Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, 7:
188–194.
Deng, J.; Lian, Y.; Wu, Z.; Xing, Y. (2013) Performance of femtosecond laser-textured cut-
ting tools deposited with WS2 solid lubricant coatings. Surface and Coatings Technology,
222: 135–143.
Deng, J.; Song, W.; Zhang, H.; Zhao, J. (2008) Performance of PVD MoS2/Zr-coated car-
bide in cutting processes. International Journal of Machine Tools and Manufacture,
48(14): 1546–1552.
Deng, B.; Tao, Y.; Guo, D. (2012) Effects of vanadium ion implantation on microstructure,
mechanical and tribological properties of TiN coatings. Applied Surface Science, 258(22):
9080–9086.
Duan, R.; Deng, J.; Ge, D.; Ai, X.; Liu, Y.; Meng, R.; Niu, J.; Wang, G. (2018) An approach
to predict derivative-chip formation in derivative cutting of micro-textured tools.
International Journal of Advanced manufacturing Technology, 95(1–4): 973–982.
Durairaj, S.; Guo, J.; Aramcharoen, A.; Castagne, S. (2018) An experimental study into the
effect of micro-textures on the performance of cutting tool. The International Journal of
Advanced Manufacturing Technology, 98: 1011–1030.
Enomoto, T.; Sugihara, T.; Yukinaga, S.; Hirose, K.; Satake, U. (2012) Highly wear-resistant
cutting tools with textured surfaces in steel cutting. CIRP Annals-Manufacturing
Technology, 61(1): 571–574.
Field, J. (2012) The mechanical and strength properties of diamond. Reports on Progress in
Physics, 75(12): 126505.
Fu, R.; Kwok, S.; Chen, P.; Yang, P.; Ngai, R.; Tian, X.; Chu, P. K. (2005) Surface modifica-
tion of cemented carbide using plasma nitriding and metal ion implantation. Surface and
Coatings Technology, 196(1–3): 150–154.
Gachot, C.; Rosenkranz, A.; Hsu, S.; Costa, H. (2017) A critical assessment of surface tex-
turing for friction and wear improvement. Wear, 372: 21–41.
Gajrani, K. K.; Suresh, S.; Sankar, M. R. (2018) Environmental friendly hard machining
performance of uncoated and MoS2 coated mechanical micro-textured tungsten carbide
cutting tools. Tribology International, 125: 141–155.
Godlevski, V.; Volkov, A.; Latyshev, V.; Maurin, L. (1997) The kinetics of lubricant pene-
tration action during machining. Lubrication Science, 9(2): 127–140.
Greenwood, J.; Williamson, J. P. (1966) Contact of nominally flat surfaces. Proceedings of
the Royal Society of London A: Mathematical, Physical and Engineering Sciences,
295(1442): 300–319.
504 Y. CHEN ET AL.

Hao, X.; Cui, W.; Li, L.; Li, H.; Khan, A. M.; He, N. (2018) Cutting performance of tex-
tured polycrystalline diamond tools with composite lyophilic/lyophobic wettabilities.
Journal of Materials Processing Technology, 260: 1–8.
Hao, S.; Zhang, Y.; Xu, Y.; Gey, N.; Grosdidier, T.; Dong, C. (2013) WC/Co composite sur-
face structure and nano graphite precipitate induced by high current pulsed electron
beam irradiation. Applied Surface Science, 285: 552–556.
Hovsepian, P. E.; Lewis, D.; M€ unz, W.-D. (2000) Recent progress in large scale manufactur-
ing of multilayer/superlattice hard coatings. Surface and Coatings Technology, 133:
166–175.
Hovsepian, P. E.; Luo, Q.; Robinson, G.; Pittman, M.; Howarth, M.; Doerwald, D.; Tietema,
R.; Sim, W.; Deeming, A.; Zeus, T. (2006) TiAlN/VN superlattice structured PVD coat-
ings: a new alternative in machining of aluminium alloys for aerospace and automotive
components. Surface and Coatings Technology, 201(1): 265–272.
Ivanov, Y. F.; Rotshtein, V.; Proskurovsky, D.; Orlov, P.; Polestchenko, K.; Ozur, G.;
Goncharenko, I. (2000) Pulsed electron-beam treatment of WC–TiC–Co hard-alloy cut-
ting tools: wear resistance and microstructural evolution. Surface and Coatings
Technology, 125(1): 251–256.
Jilek, M.; Cselle, T.; Holubar, P.; Morstein, M.; Veprek-Heijman, M.; Veprek, S. (2004)
Development of novel coating technology by vacuum arc with rotating cathodes for
industrial production of nc-(Al1xTix)N/a-Si3N4 superhard nanocomposite coatings for
dry, hard machining. Plasma Chemistry and Plasma Processing, 24(4): 493–510.
Jindal, P.; Santhanam, A.; Schleinkofer, U.; Shuster, A. (1999) Performance of PVD TiN,
TiCN, and TiAlN coated cemented carbide tools in turning. International Journal of
Refractory Metals and Hard Materials, 17(1): 163–170.
Kalss, W.; Reiter, A.; Derflinger, V.; Gey, C.; Endrino, J. (2006) Modern coatings in high
performance cutting applications. International Journal of Refractory Metals and Hard
Materials, 24(5): 399–404.
Kasuga, Y.; Yamaguchi, K. (1968) Friction and lubrication in the deformation processing of
metals: 1st Report, quantitative assessment of the surface texture of materials being
deformed under rigid tool. Bulletin of JSME, 11(44): 344–353.
Kawasegi, N.; Ozaki, K.; Morita, N.; Nishimura, K.; Yamaguchi, M. (2017) Development
and machining performance of a textured diamond cutting tool fabricated with a focused
ion beam and heat treatment. Precision Engineering-Journal of the International Societies
for Precision Engineering and Nanotechnology, 47: 311–320.
Kawasegi, N.; Sugimori, H.; Morimoto, H.; Morita, N.; Hori, I. (2009) Development of cut-
ting tools with microscale and nanoscale textures to improve frictional behavior.
Precision Engineering-Journal of the International Societies for Precision Engineering and
Nanotechnology, 33(3): 248–254.
Kim, D. M.; Lee, I.; Kim, S. K.; Kim, B. H.; Park, H. W. (2016) Influence of a micropat-
terned insert on characteristics of the tool-workpiece interface in a hard turning process.
Journal of Materials Processing Technology, 229: 160–171.
Klocke, F.; Krieg, T. (1999) Coated tools for metal cutting–features and applications. CIRP
Annals-Manufacturing Technology, 48(2): 515–525.
Koshy, P.; Tovey, J. (2011) Performance of electrical discharge textured cutting tools. Cirp
Annals-Manufacturing Technology, 60(1): 153–156.
K€ummel, J.; Braun, D.; Gibmeier, J.; Schneider, J.; Greiner, C.; Schulze, V.; Wanner, A.
(2015) Study on micro texturing of uncoated cemented carbide cutting tools for wear
improvement and built-up edge stabilisation. Journal of Materials Processing Technology,
215: 62–70.
MACHINING SCIENCE AND TECHNOLOGY 505

Lei, S.; Devarajan, S.; Chang, Z. (2009) A study of micropool lubricated cutting tool in
machining of mild steel. Journal of Materials Processing Technology, 209(3): 1612–1620.
Leyendecker, T.; Lemmer, O.; Esser, S.; Ebberink, J. (1991) The development of the PVD
coating TiAlN as a commercial coating for cutting tools. Surface and Coatings
Technology, 48(2): 175–178.
Lian, Y.; Chen, H.; Mu, C.; Deng, J.; Lei, S. (2018) Experimental investigation and mechan-
ism analysis of tungsten disulfide soft coated micro-nano textured self-lubricating dry
cutting tools. International Journal of Precision Engineering and Manufacturing-Green
Technology, 5(2): 219–230.
Lian, Y.; Deng, J.; Li, S.; Xing, Y.; Chen, Y. (2013) Preparation and cutting performance of
WS2 soft-coated tools. The International Journal of Advanced Manufacturing Technology,
67(5–8): 1027–1033.
Li, S.; Deng, J.; Zhang, G.; Zhang, K.; Zhou, Y. (2015) Dry cutting performance of tools
deposited with TiSiN–WS2/Ti–WS2 coatings. Surface Engineering, 31(12): 949–956.
Lim, S.; Lim, C.; Lee, K. (1995) The effects of machining conditions on the flank wear of
TiN-coated high speed steel tool inserts. Wear, 181: 901–912.
Ling, T.; Liu, P.; Xiong, S.; Grzina, D.; Cao, J.; Wang, Q.; Xia, Z.; Talwar, R. (2013) Surface
texturing of drill bits for adhesion reduction and tool life enhancement. Tribology
Letters, 52(1): 113–122.
Liu, Z.; An, Q.; Xu, J.; Chen, M.; Han, S. (2013) Wear performance of (nc-AlTiN)/(a-
Si3N4) coating and (nc-AlCrN)/(a-Si3N4) coating in high-speed machining of titanium
alloys under dry and minimum quantity lubrication (MQL) conditions. Wear, 305(1):
249–259.
Liu, Y.; Deng, J.; Wang, W.; Duan, R.; Meng, R.; Ge, D.; Li, X. (2018) Effect of texture
parameters on cutting performance of flank-faced textured carbide tools in dry cutting
of green Al2O3 ceramics. Ceramics International, 44(11): 13205–13217.
Liu, Y.; Deng, J.; Wu, F.; Duan, R.; Zhang, X.; Hou, Y. (2017) Wear resistance of carbide
tools with textured flank-face in dry cutting of green alumina ceramics. Wear, 372:
91–103.
Liu, Y.; Liu, J.; Du, Z. (1999) The cutting performance and wear mechanism of ceramic
cutting tools with MoS2 coating deposited by magnetron sputtering. Wear, 231(2):
285–292.
Liu, W.; Li, A.; Wu, H.; He, R.; Huang, J.; Long, Y.; Deng, X.; Wang, Q.; Wang, C.; Wu, S.
(2016) Effects of bias voltage on microstructure, mechanical properties, and wear mech-
anism of novel quaternary (Ti, Al, Zr) N coating on the surface of silicon nitride ceramic
cutting tool. Ceramics International, 42(15): 17693–17697.
Li, L.; Wang, J. (2017) Direct writing of large-area micro/nano-structural arrays on single
crystalline germanium substrates using femtosecond lasers. Applied Physics Letters,
110(25): 251901.
Lo, S. W.; Wilson, W. R. D. (1999) A theoretical model of micro-pool lubrication in metal
forming. Journal of Tribology-Transactions of the ASME, 121(4): 731–738.
Ma, S.; Prochazka, J.; Karvankova, P.; Ma, Q.; Niu, X.; Wang, X.; Ma, D.; Xu, K.; Veprek,
S. (2005) Comparative study of the tribological behaviour of superhard nanocomposite
coatings nc-TiN/a-Si3N4 with TiN. Surface and Coatings Technology, 194(1): 143–148.
Manickam, S.; Wang, J.; Huang, C. (2013) Laser–material interaction and grooving per-
formance in ultrafast laser ablation of crystalline germanium under ambient conditions.
Proceedings of the Institution of Mechanical Engineers, Part B: Journal of Engineering
Manufacture, 227(11): 1714–1723.
506 Y. CHEN ET AL.

Manory, R.; Li, C.; Fountzoulas, C.; Demaree, J.; Hirvonen, J.; Nowak, R. (1998) Effect of
nitrogen ion-implantation on the tribological properties and hardness of TiN films.
Materials Science and Engineering: A, 253(1): 319–327.
Mishra, S. K.; Ghosh, S.; Aravindan, S. (2018) Characterization and machining performance
of laser-textured chevronshaped tools coated with AlTiN and AlCrN coatings. Surface
and Coatings Technology, 334: 344–356.
M’Saoubi, R.; Ruppi, S. (2009) Wear and thermal behaviour of CVD a-Al2O3 and MTCVD
Ti(C,N) coatings during machining. CIRP Annals-Manufacturing Technology, 58(1):
57–60.
Musil, J. (2000) Hard and superhard nanocomposite coatings. Surface and Coatings
Technology, 125(1): 322–330.
Musil, J. (2012) Hard nanocomposite coatings: thermal stability, oxidation resistance and
toughness. Surface and Coatings Technology, 207: 50–65.
Obikawa, T.; Kamio, A.; Takaoka, H.; Osada, A. (2011) Micro-texture at the coated tool
face for high performance cutting. International Journal of Machine Tools &
Manufacture, 51(12): 966–972.
Prengel, H.; Santhanam, A.; Penich, R.; Jindal, P.; Wendt, K. (1997) Advanced PVD-TiAlN
coatings on carbide and cermet cutting tools. Surface and Coatings Technology, 94:
597–602.
Proskurovsky, D.; Rotshtein, V.; Ozur, G.; Markov, A.; Nazarov, D.; Shulov, V.; Ivanov,
Y. F.; Bochheit, R. (1998) Pulsed electron-beam technology for surface modification of
metallic materials. Journal of Vacuum Science & Technology A: Vacuum, Surfaces, and
Films, 16(4): 2480–2488.
Renevier, N.; Hamphire, J.; Fox, V.; Witts, J.; Allen, T.; Teer, D. (2001) Advantages of using
self-lubricating, hard, wear-resistant MoS2-based coatings. Surface and Coatings
Technology, 142: 67–77.
Renevier, N.; Lobiondo, N.; Fox, V.; Teer, D.; Hampshire, J. (2000) Performance of MoS2/
metal composite coatings used for dry machining and other industrial applications.
Surface and Coatings Technology, 123(1): 84–91.
Ruppi, S. (2005) Deposition, microstructure and properties of texture-controlled CVD
a-Al2O3 coatings. International Journal of Refractory Metals and Hard Materials, 23(4):
306–316.
Saklakoglu, I.; Saklakoglu, N.; Ceyhun, V.; Short, K.; Collins, G. (2007) The life of WC–Co
cutting tools treated by plasma immersion ion implantation. International Journal of
Machine Tools and Manufacture, 47(3): 715–719.
Sasi, R.; SubbuS., K.; Palani, I.A. (2017) Performance of laser surface textured high speed
steel cutting tool in machining of Al7075-T6 aerospace alloy. Surface and Coatings
Technology, 313: 337–346.
Schintlmeister, W.; Pacher, O.; Krall, T.; Wallgram, W.; Raine, T. (1981) Wear characteris-
tics of CVD-coated hard metal cutting tools. Powder Metallurgy International, 13(1):
26–28.
Schintlmeister, W.; Wallgram, W.; Kanz, J.; Gigl, K. (1984) Cutting tool materials coated by
chemical vapour deposition. Wear, 100(1–3): 153–169.
Shimizu, I.; Andreasen, J. L.; Bech, J. I.; Bay, N. (2001) Influence of workpiece surface top-
ography on the mechanisms of liquid lubrication in strip drawing. Journal of Tribology-
Transactions of the ASME, 123(2): 290–294.
Shum, P.; Li, K.; Shen, Y. (2005) Improvement of high-speed turning performance of
Ti–Al–N coatings by using a pretreatment of high-energy ion implantation. Surface and
Coatings Technology, 198(1): 414–419.
MACHINING SCIENCE AND TECHNOLOGY 507

Smyrnova, K.; Pogrebnjak, A.; Beresnev, V.; Litovchenko, S.; Borba-Pogrebnjak, S.;
Manokhin, A.; Klimenko, S.; Zhollybekov, B.; Kupchishin, A.; Kravchenko, Y. O. (2018)
Microstructure and physical–mechanical properties of (TiAlSiY)N nanostructured coat-
ings under different energy conditions. Metals and Materials International, 24(5):
1024–1035.
Song, J.; Huang, C.; Zou, B.; Liu, H.; Wang, J. (2012) Microstructure and mechanical prop-
erties of TiB2–TiC–WC composite ceramic tool materials. Materials & Design (1980-
2015), 36: 69–74.
Song, W.; Wang, Z.; Deng, J.; Zhou, K.; Wang, S.; Guo, Z. (2017) Cutting temperature ana-
lysis and experiment of Ti–MoS2/Zr-coated cemented carbide tool. The International
Journal of Advanced Manufacturing Technology, 93(1–4): 799–809.
Sugihara, T.; Enomoto, T. (2009) Development of a cutting tool with a nano/micro-tex-
tured surface-Improvement of anti-adhesive effect by considering the texture patterns.
Precision Engineering-Journal of the International Societies for Precision Engineering and
Nanotechnology, 33(4): 425–429.
Sugihara, T.; Enomoto, T. (2012) Improving anti-adhesion in aluminum alloy cutting by
micro stripe texture. Precision Engineering-Journal of the International Societies for
Precision Engineering and Nanotechnology, 36(2): 229–237.
Sugihara, T.; Enomoto, T. (2013) Crater and flank wear resistance of cutting tools having
micro textured surfaces. Precision Engineering-Journal of the International Societies for
Precision Engineering and Nanotechnology, 37(4): 888–896.
Sugihara, T.; Enomoto, T. (2017) Performance of cutting tools with dimple textured surfa-
ces: A comparative study of different texture patterns. Precision Engineering-Journal of
the International Societies for Precision Engineering and Nanotechnology, 49: 52–60.
Sugihara, T.; Nishimoto, Y.; Enomoto, T. (2017) Development of a novel cubic boron
nitride cutting tool with a textured flank face for high-speed machining of Inconel 718.
Precision Engineering, 48: 75–82.
Sun, J. S.; Yan, P.; Sun, X. B.; Lu, G.; Liu, F.; Ye, W.; Yang, J. Q. (1997) Tribological prop-
erties of nitrogen ion implanted WC-Co. Wear, 213(1): 131–134.
Teer, D.; Hampshire, J.; Fox, V.; Bellido-Gonzalez, V. (1997) The tribological properties of
MoS2/metal composite coatings deposited by closed field magnetron sputtering. Surface
and Coatings Technology, 94: 572–577.
Torres, C.; Heaney, P.; Sumant, A.; Hamilton, M.; Carpick, R.; Pfefferkorn, F. (2009)
Analyzing the performance of diamond-coated micro end mills. International Journal of
Machine Tools and Manufacture, 49(7–8): 599–612.
Treglio, J.; Tian, A.; Perry, A. (1993) Extending carbide tool lifetime by metal ion implant-
ation. Surface and Coatings Technology, 62(1–3): 438–442.
Uehara, Y.; Wakuda, M.; Yamauchi, Y.; Kanzaki, S.; Sakaguchi, S. (2004) Tribological prop-
erties of dimpled silicon nitride under oil lubrication. Journal of the European Ceramic
Society, 24(2): 369–373.
Uglov, V. V.; Kuleshov, A.; Soldatenko, E.; Koval, N.; Ivanov, Y. F.; Teresov, A. (2012)
Structure, phase composition and mechanical properties of hard alloy treated by intense
pulsed electron beams. Surface and Coatings Technology, 206(11): 2972–2976.
Vasumathy, D.; Meena, A. (2017) Influence of micro scale textured tools on tribological
properties at tool-chip interface in turning AISI 316 austenitic stainless steel. Wear, 376:
1747–1758.
Veprek, S.; Veprek-Heijman, M. J. (2008) Industrial applications of superhard nanocompo-
site coatings. Surface and Coatings Technology, 202(21): 5063–5073.
508 Y. CHEN ET AL.

Vorobyev, A. Y.; Guo, C. (2009) Metal pumps liquid uphill. Applied Physics Letters, 94(22):
224102.
Vorobyev, A. Y.; Guo, C. (2010) Laser turns silicon superwicking. Optics Express, 18(7):
6455–6460.
Wan, D.; Wang, J.; Mathew, P. (2011) Energy deposition and non-thermal ablation in fem-
tosecond laser grooving of silicon. Machining Science and Technology, 15(3): 263–283.
Wang, J. (2000) The effect of the multi-layer surface coating of carbide inserts on the cut-
ting forces in turning operations. Journal of Materials Processing Technology, 97(1–3):
114–119.
Wang, J.; Armarego, E. (2001) Computer-aided optimization of multiple constraint single
pass face milling operations. Machining Science and Technology, 5(1): 77–99.
Wang, X.; Kato, K.; Adachi, K.; Aizawa, K. (2003) Loads carrying capacity map for the sur-
face texture design of SiC thrust bearing sliding in water. Tribology International, 36(3):
189–197.
Wang, J.; Kuriyagawa, T.; Wei, X.; Guo, D. (2002) Optimization of cutting conditions for
single pass turning operations using a deterministic approach. International Journal of
Machine Tools and Manufacture, 42(9): 1023–1033.
Wang, Y.; Li, C.; Zhang, Y.; Yang, M.; Li B.; Dong, L.; Wang, J. (2018) Processing charac-
teristics of vegetable oil-based nanofluid MQL for grinding different workpiece materials.
International Journal of Precision Engineering and Manufacturing-Green Technology, 5(2):
327–339.
Wang, J.; Zhang, Q. (2008a) A study of high-performance plane rake faced twist drills. Part
II: predictive force models. International Journal of Machine Tools and Manufacture,
48(11): 1286–1295.
Wang, J.; Zhang, Q. (2008b) A study of high-performance plane rake faced twist drills.:
Part I: Geometrical analysis and experimental investigation. International Journal of
Machine Tools and Manufacture, 48(11): 1276–1285.
Wenzel, R. N. (1936) Resistance of solid surfaces to wetting by water. Industrial and
Engineering Chemistry, 28(8): 988–994.
Wu, Z.; Deng, J.; Chen, Y.; Xing, Y.; Zhao, J. (2012) Performance of the self-lubricating tex-
tured tools in dry cutting of Ti-6Al-4V. The International Journal of Advanced
Manufacturing Technology, 62(9): 943–951.
Xie, J.; Luo, M.; Wu, K.; Yang, L.; Li, D. (2013) Experimental study on cutting temperature
and cutting force in dry turning of titanium alloy using a non-coated micro-grooved
tool. International Journal of Machine Tools and Manufacture, 73: 25–36.
Xing, Y.; Deng, J.; Wu, Z.; Liu, L.; Huang, P.; Jiao, A. (2018) Analysis of tool-chip interface
characteristics of self-lubricating tools with nanotextures and WS2/Zr coatings in dry cut-
ting. The International Journal of Advanced Manufacturing Technology, 97: 1637–1647.
Xu, L.; Huang, C.; Liu, H.; Zou, B.; Zhu, H.; Zhao, G.; Wang, J. (2013a) Study on in-situ
synthesis of ZrB2 whiskers in ZrB2–ZrC matrix powder for ceramic cutting tools.
International Journal of Refractory Metals and Hard Materials, 37: 98–105.
Xu, Y.; Zhang, Y.; Hao, S.; Perroud, O.; Li, M.; Wang, H.; Grosdidier, T.; Dong, C. (2013b)
Surface microstructure and mechanical property of WC-6% Co hard alloy irradiated by
high current pulsed electron beam. Applied Surface Science, 279: 137–141.
Yang, M.; Li, C.; Zhang, Y.; Jia, D.; Zhang, X.; Hou, Y.; Li, R.; Wang, J. (2017) Maximum
undeformed equivalent chip thickness for ductile-brittle transition of zirconia ceramics
under different lubrication conditions. International Journal of Machine Tools and
Manufacture, 122: 55–65.
MACHINING SCIENCE AND TECHNOLOGY 509

Yashar, P. C.; Sproul, W. D. (1999) Nanometer scale multilayered hard coatings. Vacuum,
55(3–4): 179–190.
Zeng, X.; Zhang, S.; Tan, L. (2001) Multilayered (Ti, Al) ceramic coating for high-speed
machining applications. Journal of Vacuum Science & Technology A: Vacuum, Surfaces,
and Films, 19(4): 1919–1922.
Zhang, K.; Deng, J.; Ding, Z.; Guo, X.; Sun, L. (2017) Improving dry machining perform-
ance of TiAlN hard-coated tools through combined technology of femtosecond laser-
textures and WS2 soft-coatings. Journal of Manufacturing Processes, 30: 492–501.
Zhang, K.; Deng, J.; Guo, X.; Sun, L.; Lei, S. (2018) Study on the adhesion and tribological
behavior of PVD TiAlN coatings with a multi-scale textured substrate surface.
International Journal of Refractory Metals and Hard Materials, 72: 292–305.
Zhang, K.; Deng, J.; Xing, Y.; Li, S.; Gao, H. (2015) Effect of microscale texture on cutting
performance of WC/Co-based TiAlN coated tools under different lubrication conditions.
Applied Surface Science, 326: 107–118.
Zhang, D.; Shen, B.; Sun, F. (2010) Study on tribological behavior and cutting performance
of CVD diamond and DLC films on Co-cemented tungsten carbide substrates. Applied
Surface Science, 256(8): 2479–2489.
Zhao, Y.; Lu, Q.; Li, M.; Li, X. (2007) Anisotropic wetting characteristics on submicrome-
ter-scale periodic grooved surface. Langmuir, 23(11): 6212–6217.

You might also like