Vous êtes sur la page 1sur 10

Composite Structures 131 (2015) 374383

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Study on the cutting mechanics characteristics of high-strength UD-CFRP


laminates based on orthogonal cutting method
Qinglong An a,, Weiwei Ming a, Xiaojiang Cai b, Ming Chen a
a
b

School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai 200240, PR China
Shanghai Aerospace Control Technology Institute, Shanghai 201109, PR China

a r t i c l e

i n f o

Article history:
Available online 16 May 2015
Keywords:
Unidirectional carbon ber reinforced
polymer (UD-CFRP)
Orthogonal cutting
Anisotropy
Cutting force
Specic cutting energy
Machinability

a b s t r a c t
High-strength unidirectional carbon ber reinforced polymer (UD-CFRP) has gradually become one major
material for primary load-bearing structural components of aircrafts, and related machining demands are
also ever increasing. Owing to its prominent anisotropy and heterogeneity, UD-CFRP laminate has rather
poor machinability. This paper conducted orthogonal cutting tests on T700/800 high-strength UD-CFRP
laminates, and investigated machining mechanism by studying cutting mechanics characteristics in cutting process, in order to provide the basis for improving their machinability. Experimental results showed
that the cutting force and specic cutting energy of T700 and T800 UD-CFRP laminates were all signicantly directional; and at the conditions of same ber orientation angle, greater cutting force and specic
cutting energy were found for cutting of T800 UD-CFRP than that for cutting of T700. As the cutting speed
increased, main cutting force and radial thrust force both decreased and specic energy map shrank
rapidly, indicating reduced cutting energy consumption, and improved machinability of CFRPs. With
the increase in cutting depth, main cutting force and radial thrust force both exhibited increasing trends,
but specic cutting energy map narrowed, indicating improved machinability.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
As the most representative advanced resin-based composite
material, carbon ber reinforced polymer (CFRP) has been used
in the eld of aviation manufacturing since the 1970s owing to
its advantages such as high specic strength, high specic stiffness,
corrosion resistance and strong designability [1,2]. CFRP has gradually replaced traditional metal materials like aluminum alloy and
high-strength steel to become an important aviation structural
material, such as bearing fairing, empennage-level secondary
load-bearing components, fuselage, wing and other large primary
load-bearing structures [3,4]. As an example, currently, CFRP has
been fully applied to the large primary load-bearing structures of
advanced large civil aircrafts like B787, A380, A350 and A400M.
Generally, large CFRP components can be manufactured directly
through material molding. But in order to allow the CFRP components to meet the geometric dimension, shape accuracy and surface quality required for nal components, secondary machining
is often needed after the material molding, and common processing methods include edge trimming and hole machining.

Corresponding author. Tel.: +86 21 34206824; fax: +86 21 34206317.


E-mail address: qlan@sjtu.edu.cn (Q. An).
http://dx.doi.org/10.1016/j.compstruct.2015.05.035
0263-8223/ 2015 Elsevier Ltd. All rights reserved.

The characteristics of anisotropy and heterogeneity for CFRP


make its machining process differ from metal materials [5], which
also make the machining face new challenges. Effect of anisotropy
on machining of CFRP materials is mainly manifested by signicant
directionality. For example, inter-layer bonding strength of CFRP
laminates is only 520% of its tensile strength along the ber direction, which will easily lead to interlayer delamination under the
action of cutting force. Besides, during CFRP cutting process, material removal mechanism differs completely at different ber orientation angle conditions, which results in anisotropic internal acting
force inside each layer, and signicantly uctuating cutting force
overall, ultimately affecting the machining quality. Therefore,
CFRPs bring about complex material removal regime during cutting process, relatively poor machinability and easy formation of
machining defects in terms of material composition and composite
properties.
Studies on machining CFRPs mainly focus on the defects, cutting
force and tool performance when operations such as drilling,
milling and trimming are involved [611]. These machining characteristics are always related to the material removal mechanism
during the cutting process. To investigate the cutting mechanism
of CFRP, orthogonal cutting of unidirectional carbon ber reinforced polymer (UD-CFRP) laminate is often used to observe the

375

Q. An et al. / Composite Structures 131 (2015) 374383

Nomenclature
ac
CFRP
CVD
UD-CFRP
Fc
Fp
Fr

cutting depth
carbon ber reinforced polymer
chemical vapor deposition
unidirectional carbon ber reinforced polymer
main cutting force
radial thrust force
resultant cutting force

fracture and separation forms of carbon bers, elasticoplastic


deformation of resin matrices, as well as chip formation during
UD-CFRP cutting process [1214]. Fiber orientation angle h is an
intersection angle between the direction of cutting speed and the
direction of bers on non-cut layers, which is considered as the
most inuential factor in CFRP cutting process. As shown in
Fig. 1, Sreejith et al. [13] divided the material removal mechanism
during orthogonal cutting of CFRP into three categories: (1) layered
peeling fracture mechanism when the direction of cutting speed is
consistent with the ber direction; (2) extrusion shear fracture
mechanism when the direction of cutting speed is in an acute angle
with the ber direction; and (3) bending shear fracture mechanism
when the direction of cutting speed is in an obtuse angle with the
ber direction. High strength and modulus of bers, as well as very
poor cutting performance could easily lead to fracture or severe
wear of tools. As the standards-setter for aerospace grade carbon
ber materials, tensile strengths at 0 direction of Torays carbon
bers range from 3.5 Gpa for baseline T300 ones to nearly
7.0 Gpa for T800, T1000 high-strength ones. The cutting performance will differ greatly, because the strength difference between
different types of enhanced carbon bers can double.
During the cutting process, cutting force is the root of a variety
of physical phenomena. Cutting force generates cutting heat, while
the force and heat jointly inuence the machined surface quality.
Therefore, to study the cutting mechanism of a new CFRP material,
the mechanics behaviors in cutting process must be studied rst,
which is also the basis of cutting tool design and CFRP molding
for machining process. Wang et al. [14,15] systematically studied
the material removal mechanism of CFRP unidirectional laminates
based on orthogonal cutting, compares main cutting force and
radial thrust force at four typical ber orientation angles, 0/45/
90/135, through orthogonal cutting tests, and nds signicant
difference in cutting force of UD-CFRP laminates under different
directions. Cutting forces in 0 and 135 directions uctuate substantially, while those in 45 and 90 directions are relatively

(a) Layered peeling fracture

re
t
u

vc
a
c0
h

rounded edge radius


cutting width
specic cutting energy
cutting speed
clearance angle
rake angle
ber orientation angle

stable. Then, he works out that cutting force increases gradually


with the increase in ber orientation angle based on the comparison of average force. Zhang et al. [16,17] construct
two-dimensional orthogonal cutting force model of CFRP with ber
orientation angle less than 90 on the basis of orthogonal cutting
tests, as shown in Fig. 2. According to the prediction results of
the model, main cutting force basically shows a linear proportional
rise with the increase in ber orientation angle, reaching its maximum at 90 direction; radial thrust force, on the other hand,
reaches its maximum at 30 direction, then drops rapidly with
the increase in ber orientation angle, and reaches its minimum
at 90 direction. The experimental results also show that 90 is a
critical angle, beyond which severe subsurface damages will occur
and the deformation mechanisms in the cutting zone will change.
Arola et al. [18,19] studies the orthogonal cutting process of CFRP
using cutting tools with a combination of different rake and clearance angles taking into account factors such as tool parameters,
workpiece material directionality and cutting parameters.
Research results indicated the inuences of these factors on
mechanical behaviors and chip formation during cutting process,
in a descending order, as follows: direction of material itself, tool
geometry and cutting parameters. Zhang [20,21] proposed a
hypothesis that fractures of CFRP occur under the action of tool
cutting edge are all due to shearing stress in the direction perpendicular to ber exceeds the shear strength, and derived a simplied
CFRP orthogonal cutting force model under ber direction conditions. Qi et al. [22] established a force prediction model for orthogonal cutting of UD-CFRP by taking slipping, peeling, and bounding
mechanism in three different deformation areas into consideration.
Karpat et al. [23,24] proposed a mechanistic cutting force model
for milling and drilling of UD-CFRPs based on experimentally collected cutting force data using two different polycrystalline diamond cutters. The mechanistic model is also shown to be
capable of predicting cutting forces during milling and drilling of
multidirectional CFRP laminates. The experimental milling force

(b) Extrusion shear fracture

(c) Bending shear fracture

Fig. 1. Material removal mechanism of CFRP in cutting process [13].

376

Q. An et al. / Composite Structures 131 (2015) 374383

the experiment had a rake angle of 25, clearance angle of 15,


and a rounded edge radius of 15 lm, as shown in Fig. 4.
2.2. Experimental conditions

Fig. 2. A schematic of the two-dimensional orthogonal cutting of UD-CFRP [16].

measurements and predicted milling forces agree well with each


other.
In summary, the signicant directionality of mechanical behaviors during CFRP cutting process has gained the most attention.
Research ndings in the literature are mainly concentrated on
the processing along the ber direction; moreover, explanations
on cutting mechanical behaviors of CFRP under continuous variation of ber orientation angle within a 0180 range still diverge.
This paper will adopt orthogonal cutting tests to study the inuences of cutting parameters on mechanical behavioral characteristics like cutting force and specic cutting energy during UD-CFRP
material removal process.
2. Experimental procedures
Orthogonal cutting method used in this paper is a basic experimental approach commonly used in the research of cutting mechanisms of materials [25]. During cutting process, only one main
cutting edge is involved, which is perpendicular to the direction
of cutting speed. So it can simply and intuitively reect the interaction between workpiece and cutting edge.
2.1. Workpiece materials and cutting tool
Workpiece materials used in this paper were T700/LT-03A and
T800/X850 CFRP unidirectional laminates, which are isotropically
layered (all 0 direction) with 40 layers and thickness of 5 mm,
and 32 layers and thickness of 6.08 mm, respectively. Carbon ber
volume fractions of both materials were 60%; relevant mechanical
properties are shown in Table 1.
In the orthogonal cutting tests, CFRP laminates were cut into
small sheets with dimension of 25 mm  30 mm, and orthogonal
cutting tests were performed on unidirectional laminates with
ber orientation angles of 0, 15, 30, 45, 60, 75, 90, 105,
120, 135, 150 and 165, respectively, obtained by cutting 0
CFRP laminates in different directions, see Fig. 3. Chemical vapor
deposition (CVD) diamond-coated carbide cutting tool used in

Schematic diagram and test-site photo of orthogonal cutting


tests of CFRP unidirectional laminates are shown in Figs. 5 and 6,
respectively. It can be seen from Fig. 5 that the orthogonal cutting
was achieved by cutting CFRP laminates at different ber orientation angles. The experiments were completed on a
KENT-KGS-1020AH surface grinder, where the grinding wheel
was replaced by a specially designed ywheel installed with cutting tool. The working table of surface grinder remained xed,
and orthogonal cutting motion was accomplished by spindle rotation; besides, cutting depth was achieved by Z direction feed
motion. Cutting forces were collected with a Kistler-9272
dynamometer, the charge signals obtained by the dynamometer
were amplied through Kistler-5017B amplier; nally, acquisition of cutting forces was completed using computer signal acquisition system. Single-factor experimental method was adopted.
Cutting speed vc and cutting depth ac both had ve levels, which
were 100, 150, 200, 250 and 300 m/min, and 0.005, 0.01, 0.015,
0.020 and 0.025 mm, respectively.
3. Results and discussion
3.1. Effects of ber type and ber orientation angle on cutting force
Fig. 7 shows the effects of ber type and ber orientation angle
h on cutting force under conditions of vc = 200 m/min, ac = 20 lm.
As can be seen from the gure, main cutting force Fc and radial
thrust force Fp both have signicant uctuations under the inuences of different ber types and ber orientation angles. There
exist the largest values with h = 90 for both Fc and Fp, which almost
shows the same as Zhang observations [26]. It also can be seen
from Fig. 7 that the inuence of ber type on cutting force was relatively stable and direct. T800, whose ber strength was higher,
had Fc and Fp both about 1030 N greater than T700 at the same
ber orientation angle.
As can be seen from Fig. 7(a), when ber orientation angle h
changes between 0 and 75, main cutting force Fc basically
increases linearly with the increase of h. From 75 to 90, Fc soared
quickly. After reaching the peak at 90, Fc quickly fell to a lower
level as h became larger, except for a small peak in the 165 direction, Fc uctuated rather slightly within a 105180 range. It can
thus be seen that in the forward ber direction of h < 90, Fc showed
linear growth as h increased, the smaller the forward ber orientation angle, the smaller the value of Fc, whereas the larger the forward ber orientation angle, the greater the value of Fc. In the
reserve ber direction of h > 90, Fc was maintained at a relatively
low level, which did not change markedly with changes in ber orientation angle, except for a slight rise in the direction of h = 165.
It can be seen from the above analysis that the effect of ber orientation angle h on main cutting force Fc exhibited a strong anisotropy. Fc of T700 and T800 CFRPs both presented a signicantly
soaring peak at h = 90, which was, Fc was the largest at a vertical

Table 1
Mechanical properties of unidirectional CFRP laminates.
Material

Tensile strength
(MPa)

Tensile modulus
(GPa)

Compressive
strength (MPa)

Compressive
modulus (GPa)

In-plane shear
strength (MPa)

90 tensile strength
(MPa)

Bending strength
(MPa)

T700
T800

2450
2840

125
168

1430
1570

145

92
98

70
80

1580
1670

Note: Unspecied mechanical property parameters all represent the properties in 0 direction.

Q. An et al. / Composite Structures 131 (2015) 374383

377

Fig. 3. Workpiece materials (0 CFRP laminates).

a 030 range, and slower within a 3090 range, which peaked


in the vertical direction of h = 90. Afterwards, as h continued to
grow into the reserve ber direction of h > 90, Fp showed a significant decline, which further decreased with the increase of h.
Similar to Fc, Fp also had a slight rise in the direction of h = 165.
It can be seen from the above analysis that radial thrust force Fp
also had signicant directionality, but it was somewhat different
from Fc. Because UD-CFRP laminate was a material with very severe bouncing-back of machined surfaces, Fp was always greater
than Fc at the same ber orientation angle [27]. Moreover,
bouncing-back resilience differed distinctively in different directions. Maximum of Fp also appeared in the direction of h = 90,
reaching 148.8 N and 119.7 N for T800 and T700, respectively;
however, the peaks did not soar signicantly as Fc but were relatively stable. Minimum values of Fp appeared only in the parallel
direction of h = 0, which were 55.3 N for T800 and 42.9 N for
T700. This is because the bers have the highest longitudinal
strength, which will lead to a high stiffness in the
parallel-to-ber direction, and hence the resistance to the vertical
compression of the cutting tool becomes higher when h changes to
90. Meanwhile, the material was removed by shearing fracture of
bers, and there will be largest resistance when cutting a bundle of
well-bonded bers with h = 90 [26].
Fig. 4. Orthogonal cutting tool in experiments.

cutting direction, which could reach 96.3 N and 72.9 N, respectively, for T800 and T700. The position with the best machinability
was the reverse ber direction of 120150, where Fc values of
T800 and T700 were the lowest, which were 27.2 N and 17.9 N,
respectively. In addition, the value of Fc was also relatively low
when the angle was 0 or 180, which was 32.3 N for T800, and
26.7 N for T700, respectively.
As to the radial thrust force Fp, it can be seen from Fig. 7(b) that
in the forward ber direction with h < 90, Fp showed piecewise
linear growth as angle h increased; the growth was faster within

3.2. Effects of ber type and ber orientation angle on specic cutting
energy and specic energy map
Specic cutting energy refers to the energy consumed for
removal of unit volume of material in orthogonal cutting [25],
which can be expressed as:

Fc
tac

where Fc is the main cutting force, and t and ac are cutting width
and cutting depth in orthogonal cutting, respectively. Specic cutting energy is an important parameter for machinability of

378

Q. An et al. / Composite Structures 131 (2015) 374383

Dynamic cutting force acquisition system

Spindle

Flywheel
Data acquisition and analysis
Cutter
CFRP laminate
Kistler dynamometer

Amplifier

Fig. 5. Scheme of orthogonal cutting of CFRP unidirectional laminates.

0
Cutter

vc
ac

Workpiece

Fc

Fp
Fig. 6. Photo of orthogonal cutting of CFRP unidirectional laminates.

materials, which directly reects the difculty of chip separation in


cutting.
Fig. 8 shows the specic cutting energy u and resultant cutting
force Fr of T700 and T800 CFRPs. As can be seen, T800 performed a
poorer machinability than T700; and under the same cutting conditions, T800 CFRP consumed more energy. Machinability of both
CFRPs exhibited signicant directionality. With changes in the ber
orientation angle, energy consumed during machining differed distinctly. There existed a critical angle of h = 90, beyond which severe subsurface damages will occur and the deformation
mechanisms in the cutting zone will change. The removal of ber
material in UD-CFRP cutting was mainly completed by shear fracture and extrusion fracture in 90 ber orientation. There would be
a larger resistance when cutting a bundle of well-bonded bers
with h = 90. Therefore, machinability was the poorest, and chip
separation was most difcult in the vertical direction of h = 90.
In the forward ber direction of h < 90, machinability became
poorer linearly with the increase in h. However, in the reverse ber
direction of h > 90, machinability was superior to that in the forward ber direction, and chip separation became easier. The parallel direction of h = 0 (or 180) was also a direction where material
removal was rather easily achieved.
In the aeronautical application of CFRP, drilling is the most common processing method of secondary machining. This paper would
study specic cutting energy during drilling process based on

orthogonal cutting test combining CFRP unidirectional laminate


drilling principle by establishing cutting specic energy map.
Specic steps were: if drill bit made a 360 circumferential rotary
motion, the intersection angle between drill edge and direction of
bers in a certain layer during cutting would also change within
the range of 0360. Taking into account that all of the intersection
angles between cutting direction of tool and ber direction could
be described by ber orientation angle h, intersection angles within
a 0180 range could be described directly by ber orientation
angle h, while the intersection angles ranging 180360 could be
described by ber orientation angle h after subtracting 180.
Specic energy map of T700 and T800 unidirectional CFRP laminates during circumferential cutting can be obtained based on
the test results in Fig. 8, as shown in Fig. 9. As can be seen from
the gure, the specic energy of monolayer CFRP laminate is significantly directional within a single cutting cycle. The entire map
was in a pigeon-shape: specic cutting energy was smaller in
the second and fourth quadrants where cutting was performed in
the reverse ber direction, and larger in the rst and third quadrants where cutting was performed in the forward ber direction;
maximum specic energy appeared when cutting was performed
in the perpendicular to ber direction (h = 90/270), which was
about 3 times that of the minimum value, forming the wings of
the pigeon-shaped map; there were slight specic energy
increases in the positions close to the horizontal axis

Q. An et al. / Composite Structures 131 (2015) 374383

379

3.3. Effects of cutting speed on cutting force

Fig. 7. Effects of ber type and ber orientation angle on cutting force in horizontal
direction (a) and vertical direction (b).

(h = 165/345) in the second and fourth quadrants, which formed


the head and tail of the pigeon-shaped map. The pigeon shape
in the specic energy map of monolayer CFRP laminate depended
on the properties of CFRP material itself, including the fracture
property of carbon bers, cutting performance of epoxy resins,
and anisotropy of long-ber-reinforced composites. The shapes of
specic energy maps for monolayer T700 and T800 CFRP laminates
of different ber types were basically the same.

Fig. 8. Effects of ber type and ber orientation angle on specic energy and
resultant cutting force.

Fig. 10 shows the effects of cutting speed on cutting force for


T700 CFRP. Overall, with an increasing cutting speed, main cutting
force Fc and radial thrust force Fp both decreased to some extent, Fc
was rather obviously affected by vc, while Fp was relatively insensitive to vc, and exhibited mild overall changes.
Within the test range of cutting parameters, the inuence of
cutting speed vc on main cutting force Fc was divided into two
stages: (1) within the vc range of 100200 m/min, Fc declined
rapidly with changes in vc; increased cutting speed signicantly
facilitated the cutting, and cutting of carbon bers also became
more smooth; (2) within the vc range of 200300 m/min, Fc no
longer changed obviously with changes in vc, and showed a slight
upward trend, that was, after the cutting speed exceeded
200 m/min, increased cutting speed was longer able to further
reduce the main cutting force Fc. It was also investigated by
Sreejith et al. [13] that the range of 200300 m/min of cutting
speed is the most suited for the machining of CFRP based on the
steady specic cutting pressure criterion.
Taking into account different ber orientation angles, main cutting force Fc in these two stages also had regularities: (1) within the
vc range of 100200 m/min, the anisotropy of Fc was more prominent at different ber orientation angles, which was higher for
h = 45/90 than h = 0/135/165; (2) within the vc range of 200
300 m/min, although Fc maintained the anisotropy, difference of
Fc at different ber orientation angles was no longer prominent,
that was, under high-speed cutting conditions, anisotropy of
CFRP had a gradually weakening trend. This was because high cutting speed would lead to cutting off the bers easily. The high cutting temperature caused by high-speed cutting would also result in
softening the resin base material and reducing the cutting forces.
These all helped to weaken the anisotropy of UD-CFRP.
In contrast, cutting speed vc had smaller inuence on radial
thrust force Fp. In the directions of h = 45/90 with larger Fp, Fp
exhibited a downward trend with the increase of vc, but attened
after 200 m/min. In the directions of h = 0/135/165, the inuence
of vc was rather small. Fp maintained a stable level under various
cutting speed conditions, except at the minimum cutting speed
vc = 100 m/min, where the value of Fp was slightly larger. Radial
thrust force Fp was insensitive to changes of cutting speed vc in
the reverse ber and parallel to ber directions. In addition, the
anisotropy of CFRP also showed a gradually weakening trend under
high-speed cutting conditions according to the changing regularity
of thrust force.
3.4. Effects of cutting depth on cutting force
Fig. 11 shows the effects of cutting depth on cutting force for
T700 CFRP. Overall, with the increasing of cutting depth ac, main
cutting force Fc and radial thrust force Fp both showed upward
trends, which were directly related to the growth in volume of cutting materials.
Under different ber orientation angle h conditions, the regularity of inuence of ac on cutting force was also slightly different. In
the ve ber directions shown in Fig. 11(a), main cutting force Fc
all exhibited an increasing and attening trend (except for h
= 90 direction) after cutting depth ac was increased to 15 lm,
which can be considered as a threshold. When ac was smaller than
this value, Fc grew faster as ac became larger; while after ac
exceeded this threshold, the growth of Fc signicantly slowed
down. It was because the cutting tool had a rounded edge radius
of re = 15 lm, and there was a size effect for the inuence of cutting
depth on main cutting force during UD-CFRP cutting process [28].
The critical value of ac was close to the rounded edge radius re.
When ac < re, Fc was more sensitive to the changes in ac, and grew

380

Q. An et al. / Composite Structures 131 (2015) 374383

Fig. 9. Specic energy map in cutting T700/T800 CFRP of one uniform unidirectional laminate.

Fig. 10. Effects of cutting speed on cutting force in horizontal direction (a) and vertical direction (b).

Fig. 11. Effects of cutting depth on cutting force in horizontal direction (a) and vertical direction (b).

Q. An et al. / Composite Structures 131 (2015) 374383

faster with increase in ac; when ac > re, Fc had reduced sensitivity to
changes in ac, and grew slower with an increasing ac. Size effect
had certain directionality, which appeared in the directions of
h = 0/45/135/165, but was not obvious in h = 90 direction. In
addition, directional difference of main cutting force Fc did not
change with ac, and the difference in Fc between different ber
directions almost had no change (except for h = 90) under different cutting depths.
The effects of cutting depth ac on radial thrust force Fp, in contrast, was relatively simple, without size effect near the rounded
edge radius re. Fp roughly showed a proportional growth trend with
the increasing ac in different ber directions. Meanwhile, the anisotropy of Fp was also amplied with the increase of ac. When cutting UD-CFRP, it would result in more pronounced directionality of
Fp with increasing ac. Maximum value of Fp occurred in the h = 90
direction, while the minimum value occurred in the h = 0 direction. The difference between maximum and minimum would be
amplied in proportion with the increase of ac.
3.5. Effects of cutting speed on specic energy map
Fig. 12 shows the maps of specic energy for T700 CFRP at different cutting speeds. As can be seen from the gure, cutting speed

381

vc has rather signicant effect on specic cutting energy u; with the


acceleration of vc, the specic energy map shrank quickly, indicating reduced cutting energy consumption, and improved machinability of CFRP. As can be seen from Fig. 12, specic energy map
all maintained a pigeon-like shape within the range of vc
= 100200 m/min; the pigeon area decreased in the same proportion as vc increased. This indicated that the inuence of cutting
speed vc on the machinability of CFRP in various ber directions
was uniform at low cutting speeds; besides, cutting speed had little
inuence on the anisotropy of CFRP. However, at the high-speed
range of vc = 200300 m/min, the shape of pigeon-like map
began to change. With an increasing of vc, map area had an amplifying trend, which indicated that at the high-speed range, the
machinability of CFRP in various ber directions gradually became
consistent, and anisotropy was gradually weakened during cutting
process. It shows a similar trend with the effect of vc on the cutting
forces. On one hand, high cutting speed would lead to cutting off
the bers easily, on the other hand, high-speed cutting would
result in high cutting temperature that will soften the resin base
material and help reduce the overall chip separation energy. But
there existed a critical temperature for stable machining when less
specic energy was consumed [13]. It may lead to the little difference of specic energy map between 100 m/min and 200 m/min.

Fig. 12. Effects of cutting speed on specic energy map of T700 CFRP.

382

Q. An et al. / Composite Structures 131 (2015) 374383

Fig. 13. Effects of cutting depth on specic energy map of T700 CFRP.

3.6. Effects of cutting depth on specic energy map

4. Conclusions

Fig. 13 shows the maps of specic cutting energy for T700


CFRP under different cutting depth conditions. As can be seen
from the gure, the shape and numerical range of specic energy
map shrunk with the increase of ac. This was also observed by
Sreejith et al. [13]. At the minimum cutting depth of ac = 5 lm,
the specic energy map was most ample, and the specic
cutting energy u in various ber directions was the highest,
showing relatively strong difculty in machinability. It is because
the tested T700 carbon ber had a lament diameter of about 6
8 lm, and cutting may occur inside the bers with ac = 5 lm,
which would consume more energy. Within a cutting depth
range of ac = 1025 lm, specic energy map roughly maintained
the relatively regular pigeon-like shape. It could be described
as follows: in the forward ber direction of h < 90 and the
reverse ber direction of h > 90, specic energy map narrowed
as ac increased, indicating improved machinability; in a h = 90
direction, specic energy was less affected by ac, which maintained stable after ac exceeded the ber diameter, indicating that
the machinability in this direction is basically not affected by the
cutting depth.

This paper studied the effects of ber orientation angle on cutting force characteristics of UD-CFRP laminates using the orthogonal cutting test, and reached the following conclusions.
Cutting force and specic cutting energy of T700 and T800
CFRPs were all signicantly directional. At the same ber orientation angle conditions, T800 always had greater cutting force and
specic cutting energy than T700.
Fiber orientation angle had signicant effects on main cutting
force Fc and radial thrust force Fp. In the forward ber direction
of h < 90, Fc and Fp increased linearly as the angle h increased; in
the reverse ber direction of h > 90, Fc and Fp maintained at relatively low levels except for a slight rise in the h = 165 direction.
Fc and Fp had the peak value in the vertical direction of h = 90.
As the cutting speed increased, Fc and Fp both decreased; under
high-speed cutting conditions, the cutting anisotropy of CFRP
showed a weakening trend. With the increase of cutting depth, Fc
and Fp both exhibited upward trends; Fc has a size effect when cutting depth is close to the rounded edge radius.
With the increase of cutting speed, specic energy map shrank
quickly, indicating reduced cutting energy consumption, and

Q. An et al. / Composite Structures 131 (2015) 374383

improved machinability of CFRPs. Specic cutting energy was less


affected by ac in h = 90 direction, while in the forward ber direction of h < 90 and reverse ber direction of h > 90, specic cutting
energy would decrease with an increasing of ac. Specic energy
map narrowed, which indicated improved machinability.
According to the above analysis about cutting mechanics characteristics of UD-CFRP, a suitable high cutting speed with appropriate large cutting depth is helpful for stable machining of CFRP. To
get a comprehensive study on the cutting mechanism of UD-CFRP,
more experiments should be carried out in terms of tool parameters, cutting temperature and machining quality. In the future
work, an optimal model will be built and used in design of cutting
tool and improvement of machining quality.
Acknowledgement
This research is supported by National Natural Science
Foundation of China (Nos. 51475298 & 51105253).
References
[1] Komanduri R. Machining of ber-reinforced composites. Mach Sci Technol
1997;1(1):11352.
[2] Teti R. Machining of composite materials. CIRP Ann Manuf Technol
2002;51(2):61134.
[3] Soutis C. Fibre reinforced composites in aircraft construction. Prog Aerosp Sci
2005;41(2):14351.
[4] Chen SJ. Composite technology and large aircraft. Acta Aeronaut Astronaut Sin
2008;3:60510.
[5] Knig W, Wulf C, Gra P, Willerscheid H. Machining of bre reinforced plastics.
CIRP Ann Manuf Technol 1985;34(2):53748.
[6] Colligan K, Ramulu M. The effect of edge trimming on composite surface plies.
Manuf Rev 1992;5(4):27483.
[7] Davim JP, Reis P. Study of delamination in drilling carbon ber reinforced
plastics (CFRP) using design experiments. Compos Struct 2003;59(4):4817.
[8] Hosokawa A, Hirose N, Ueda T, Furumoto T. High-quality machining of CFRP
with high helix end mill. CIRP Ann Manuf Technol 2014;63(1):8992.
[9] Xu JY, An QL, Chen M. A comparative evaluation of polycrystalline diamond
drills in drilling high-strength T800S/250F CFRP. Compos Struct 2014;117:
7182.

383

[10] Duro LMP, Tavares JMR, de Albuquerque VHC, Gonalves DJ. Damage
evaluation of drilled carbon/epoxy laminates based on area assessment
methods. Compos Struct 2013;96:57683.
[11] Tsao CC, Hocheng H, Chen YC. Delamination reduction in drilling composite
materials by active backup force. CIRP Ann Manuf Technol 2012;61(1):914.
[12] Koplev A, Lystrup A, Vorm T. The cutting process, chips, and cutting forces in
machining CFRP. Composites 1983;14(4):3716.
[13] Sreejith PS, Krishnamurthy R, Malhotra SK, Narayanasamy K. Evaluation of
PCD tool performance during machining of carbon/phenolic ablative
composites. J Mater Process Technol 2000;104(1):538.
[14] Wang DH, Ramulu M, Arola D. Orthogonal cutting mechanisms of graphite/
epoxy composite. Part I: unidirectional laminate. Int J Mach Tools Manuf
1995;35(12):162338.
[15] Wang DH, Ramulu M, Arola D. Orthogonal cutting mechanisms of graphite/
epoxy composite. Part II: multi-directional laminate. Int J Mach Tools Manuf
1995;35(12):163948.
[16] Zhang LC, Zhang HJ, Wang XM. A force prediction model for cutting
unidirectional bre-reinforced plastics. Mach Sci Technol 2001;5(3):293305.
[17] Zhang LC. Cutting composites: a discussion on mechanics modelling. J Mater
Process Technol 2009;209(9):454852.
[18] Arola D, Ramulu M, Wang DH. Chip formation in orthogonal trimming of
graphite/epoxy composite. Composites Part A 1996;27(2):12133.
[19] Arola D, Ramulu M. Orthogonal cutting of ber-reinforced composites: a nite
element analysis. Int J Mech Sci 1997;39(5):597613.
[20] Zhang HJ. Study on the drilling technology of CFRP [Ph.D. thesis]. Beijing:
Beijing University of Aeronautics and Astronautics; 1998.
[21] Zhang HJ. Study on cutting forces of unidirectional carbon ber reinforced
plastics under orthogonal cutting. Acta Aeronaut Astronaut Sin 2005;5:6049.
[22] Qi Z, Zhang K, Cheng H, Wang D, Meng Q. Microscopic mechanism based force
prediction in orthogonal cutting of unidirectional CFRP. Int J Adv Manuf Tech
2015:111.
[23] Karpat Y, Bahtiyar O, Deger B. Mechanistic force modeling for milling of
unidirectional carbon ber reinforced polymer laminates. Int J Mach Tools
Manuf 2012;56:7993.
[24] Karpat Y, Bahtiyar O, Deger B, Kaftanoglu B. A mechanistic approach to
investigate drilling of UD-CFRP laminates with PCD drills. CIRP Ann Manuf
Technol 2014;63(1):814.
[25] Shaw MC. Metal cutting principles. New York: Oxford University Press; 2005.
[26] Hu NS, Zhang LC. Some observations in grinding unidirectional carbon brereinforced plastics. J Mater Process Technol 2004;152(3):3338.
[27] Wang XM, Zhang LC. An experimental investigation into the orthogonal
cutting of unidirectional bre reinforced plastics. Int J Mach Tools Manuf
2003;43(10):101522.
[28] Caprino G, Santo L, Nele L. Interpretation of size effect in orthogonal machining
of composite materials. Part I: unidirectional glass-bre-reinforced plastics.
Composites Part A 1998;29(8):88792.

Vous aimerez peut-être aussi