Вы находитесь на странице: 1из 11

Measurement 69 (2015) 2030

Contents lists available at ScienceDirect

Measurement
journal homepage: www.elsevier.com/locate/measurement

A model for predicting surface roughness in single-point


diamond turning
Junyun Chen a,, Qingliang Zhao b
a
b

College of Vehicles and Energy, Yanshan University, Qinhuangdao 150001, China


Center for Precision Engineering (CPE), Harbin Institute of Technology, Harbin 150001, China

a r t i c l e

i n f o

Article history:
Received 9 June 2014
Received in revised form 16 October 2014
Accepted 3 March 2015
Available online 14 March 2015
Keywords:
Prediction model
Relative tool-work vibration
Swelling effect
Surface roughness
Single-point diamond turning

a b s t r a c t
The relative tool-work vibration is not generalized enough to represent the actual displacement between tool and workpiece in previous prediction models. This is due to the fact
that the vibration was assumed as a steady simple harmonic motion and was only measured before turning process. In this study, an improved method is presented to evaluate
the actual relative tool-work vibration. By using this method the vibration information
obtained is more credible, as it contains the components caused by machine tool error,
cutting force, material property and changing of cutting parameters. Moreover, the swelling effect is analyzed using a new evaluating method and taken into account for predicting
surface roughness. On the basis of analyzing both the relative vibration and the swelling
effect, a model is proposed for predicting surface roughness Ra in single point diamond
turning. Prediction results prove that this model is a closer approximation of the actual
turning process as compared to the previous models and shows a higher predicting
accuracy of surface roughness.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Single point diamond turning (SPDT) is a promising tool
based machining technology, which can be used for manufacturing optical components and precision molds. The
main feature of SPDT is its ability to produce high-quality
surface nish on the order of nanometers while meeting
tight form tolerances on the order of micrometers [1]. For
products, surface roughness is an important index of
product quality and technical requirement [2]. In some
cases, the surface roughness is required to be kept within
a certain range rather than the possible lowest value,
especially in the case of precision mold inserts of optical
parts used for injection molding process. On the other
hand, SPDT is a complicated process inuenced easily by
Corresponding author at: No. 438, Hebei Avenue, Qinhuangdao, Hebei
Province, China. Tel.: +86 15227248304.
E-mail address: sophiacjy@ysu.edu.cn (J. Chen).
http://dx.doi.org/10.1016/j.measurement.2015.03.004
0263-2241/ 2015 Elsevier Ltd. All rights reserved.

the material swelling effect because of a ne feed rate


and high spindle speed [3]. Consequently, the investigation
on prediction of surface roughness in SPDT is signicant
and necessary in order to control the desired surface
roughness of product in a fast and effective manner.
Many researchers are interested in the prediction of
surface roughness and research in this eld has yielded
some useful ndings along with successful experience
through the use of approaches based on machining theory,
experimental investigation and articial intelligence
[2,48]. Each approach possesses its own unique advantages and disadvantages. However, machining theory
based approach together with the experimental investigation approach appears to be the most promising approach.
Combining the two approaches allows the accurate prediction of surface roughness along with this approach aiding
with the evaluation and improvement of machine tool
performance.

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

In machining theory based approach, a model based on


the theory of machining is used to simulate the creation of
the machined surface prole and visualize the surface
topography along with assessing surface roughness [2].
The surface roughness prole is generated with the repetition of the tool tip prole at intervals of feed per revolution (ideal surface roughness prole) plus a relative
displacement between the tool and workpiece. Sata et al.
[9] found that the roughness prole was composed of tool
feed component, swelling of workpiece, spindle rotation
error and chatter vibration error using spectrum analysis
method. Takasu et al. [10] estimated the surface roughness
prole as a function of both the ratio of vibration amplitude to geometrical roughness and the phase shift of the
vibration to work rotation. He established a theoretical
model for creation of roughness prole and also indicated
that, due to the tool interference, surface roughness in
the tool feed direction can be made much ner than the
sum of whole vibration amplitude and geometrical roughness. On the basis of the model created by Takasu et al.,
Cheung et al. [11] established a three-dimensional surface
topography simulation model, which takes into account
the tool geography, the machining condition as well as
the relative tool-work vibration in the kinematics of diamond turning process. The surface topography was generated by a linear mapping of the predicted surface
roughness proles on the surface elements of a cross lattice. Lee et al. [12] presented a dynamic surface topography
model for the prediction of nano-surface generation in
which an additional displacement caused by material
induced vibration was introduced into the model as compared to the previous model.
2. Previous theoretical models
The basic idea for the theoretical models mentioned
above is the fact that the actual surface roughness prole
is formed by an ideal surface roughness prole plus a relative displacement between the tool and workpiece, which
is achieved by theoretic calculation. Ideal surface roughness prole is determined by cutting conditions, while a
relative displacement between tool and workpiece was
considered as the relative tool-work vibration. In the case
of the primary models, materials were assumed as
homogeneous and isotropic and the relative vibration
was assumed as a steady simple harmonic motion. For anisotropy of crystalline materials, an enhanced model, which
adds material induced vibration to the primary model, was
used. Although, they are able to predict surface roughness
with low error, there are two major issues impacting prediction accuracy in previous theoretical models. First, the
relative tool-work vibration was assumed as invariable,
and it was measured before turning process. The second
is related to the material swelling effect, which was
ignored in previous models.
In the cutting process, the actual relative tool-work
vibration was caused by machine tool error, cutting force,
material property and change in cutting parameters.
Some scholars have tried to analyze the relative vibration
using the spectrum method or by measuring the force in
cutting process [9,1315]. However, no research work

21

has been reported which takes the actual relative toolwork vibration into account when establishing a model.
On the other hand, according to the results of roughness
prole based on the spectrum analysis, it can be concluded
that the material swelling is an important part contributing to the surface roughness prole [9,13]. Additionally,
empirical data suggests and supports that material swelling obviously changes the surface roughness prole [3].
However, no report predicts surface roughness while considering the effect of material swelling.
In the present study, a prediction model is presented to
predict the surface roughness in the SPDT process, which
takes both actual tool-work vibration and material swelling into account. It is almost impossible to measure the
actual tool-work vibration directly in cutting process. To
overcome this challenge, a concept of equivalent amplitude was proposed to aid with the evaluation of the actual
tool-work vibration with the assistance of experiments.
Furthermore, the swelling proportion of every material
was dened to quantify the swelling effect, and the relation between the swelling effect and cutting parameters
was investigated by means of experiments.
3. Experimental setup
A series of face cutting tests were conducted on a fouraxis CNC ultra-precision machine tool (made by NachiFujikoshi Corp., Japan) shown in Fig. 1 (left). A diamond
tool used in tests is shown in Fig. 1 (right), with a rake
angle of 0, a front clearance angle of 6 and a tool-nose
radius of 0.5 mm.
The tests were carried out on three kinds of materials
including copper (Cu), aluminum alloy (Al7075-T6) and
electroless-nickel (NiP) during studying relative tool-work
vibration and the swelling effect. Aluminum alloy and copper were available in market, while samples of NiP were
prepared on an aluminum alloy rod (7075-T6). In order to
achieve both the required hardness and good machinability,
the compounding of coating solution was optimized to generate medium-phosphorus NiP, which possesses a coating
depth of more than 50 lm and a hardness of 50HRC.
Table 1 tabulates the cutting conditions in the tests studying relative tool-work vibration and the swelling effect.
The surface prole was measured about 10 mm in
length by contact probe prolometer, Form Talysurf PGI
1240 (Taylor Hobson Ltd.) in 2D, while the surface topography was measured by a non-contact type surface measurement system, White Light Interferometer Veeco NT1100
(WLI, Veeco Metrology Group) in 3D for each sample. The
measurement data was then processed with MATLAB software. The diamond tool wear was observed by a scanning
electronic microscope (SEM, Hitachi S-4700) and an optical
measuring microscope (STM6, Olympus, Japan).
4. The relative tool-work vibration
4.1. Evaluating the relative tool-work vibration
Relative vibration may be caused by machine tool error,
cutting force, material properties and change in cutting

22

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

Fig. 1. CNC ultra-precision machine tool (left) and diamond tools used in tests (right).

Table 1
Cutting conditions for experiments.
Term
No.

Spindle speed
(rpm)

Feed rate
(mm/min)

Depth of cut
(lm)

Tool-nose
radius (mm)

Term
no.

Spindle
speed (rpm)

Feed rate
(mm/min)

Depth of cut
(lm)

Tool-nose
radius (mm)

1
2
3
4
5
6
7
8
9
10
11

1000
1000
1000
1000
1000
1000
1000
1500
1500
1500
1500

25
30
35
40
40
40
40
25
30
35
40

2
2
2
2
4
6
8
2
2
2
2

0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5

12
13
14
15
16
17
18
19
20
21

1500
1500
1500
2000
2000
2000
2000
2000
2000
2000

40
40
40
25
30
35
40
40
40
40

4
6
8
2
2
2
2
4
6
8

0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5

parameters during the cutting process. The relative vibration between the tool and workpiece translates onto the
machined surface and is very difcult to be measured
directly. Therefore it is feasible and applicable to extract
the vibration information from the machined surface.
After which, the relationship between the relative vibration and each corresponding factor can be analyzed to predict the surface roughness.
Previous models consider the amplitude of basic frequency of spindle to be the amplitude of simple harmonic
motion, which indicates that the amplitude of relative
vibration rather than frequency has a dominant impact
on surface roughness. As a result, considerable attention
needs to be given to the amplitude of actual tool-work
vibration, and then a denition of equivalent amplitude
was proposed to evaluate it.
A part of measured surface prole was shown in Fig. 2.
The location of diamond tool tip xi ; Y i xi can be extracted
from the measured surface roughness prole to form the
tool locus in radial direction, which represents the displacement between the tool and workpiece, i.e., the relative tool-work vibration.
By taking X-axis along the tool feed and Y-axis along the
infeed cutting direction, the number of cutting edges
duplicated on the machined surface along radial direction
is given as

N dL  x1 =se

where L is the length of measured surface roughness prole, s is tool feed per work revolution, and de means

rounding down to the nearest whole unit. In the measured


surface roughness prole, the location of every point forming tool locus can be extracted according to X value, as
follows:

xi x1 i  1Dx x1 i  1s

with i 1; 2; . . . ; N. Consequently, the location of every


point xi ; Y i xi can be achieved from the measured roughness prole, as shown in Fig. 2. The tool locus, i.e. the actual
relative tool-work vibration can be dened as

Y t xi Y i xi  minfyi xi g

with i 1; 2; . . . ; N. This vibration will increase the surface


roughness of machined surface. In order to evaluate the
vibration easily, the actual relative tool-work vibration
was simplied as a simple harmonic motion. The principle
of simplication lies on an assumption that both the relative tool-work vibration and simple harmonic motion have
the same impact on surface roughness. That is, both the
prole of tool locus and the curve of simple harmonic
motion were regarded as surface roughness proles and
had the same arithmetic roughness value Ra. As a result,
the amplitude of simple harmonic motion is dened as
equivalent amplitude, which represents the degree of
changing surface roughness with respect to the actual relative tool-work vibration. The simplication procedure for
the vibration is described below.
Assuming a simple harmonic motion as

  cos2pfx
Y h x A1

23

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

Micrometers

Measured surface profile

x Y x

( xi , Yi ( xi ))

( x2 , Y2 ( x2 ))

Tool locus
Millimeters

Fig. 2. Extract tool locus from the measured surface roughness prole.

and its discretization form can be expressed as

  cos2pf Dxi  1
Y h xi A1

where Dx 1=mf and f was assumed as 1=2s, m is an optimized positive integer, i 1; 2; . . . ; N 1 , N 1 dNs=Dxe. The
arithmetic roughness value Rat calculated from the prole
of tool locus and the arithmetic roughness value Rah calculated from the curve of simple harmonic motion can be
given as Eqs. (6) and (7) respectively.



PN

N 
1X

i1 Y t xi 
Rat

Y t xi 

N i1 
N
Rah

N1 

1 X


Y h xi  Y h xi 
N1 i1

Let Eq. (6) equal Eq. (7), then the equivalent amplitude
 can be calculated as
A


P 
P

N21 Ni1 NY t xi  Ni1 Y t xi 
 P 

A P  

pDxi1
1
1
N2 Ni1
1  cos pDxi1
 Ni1
N 1 1  cos s

s
8
4.2. Effect of cutting parameters
When depth of cut is set as 2 lm, the actual relative
tool-work vibration as a function of the change of feed rate
is shown in Fig. 3 at different spindle speeds. It can be seen
that the equivalent amplitude has no obvious uctuation
except for Al7075 under spindle speed of 2000 r/min,
which may be caused by the inhomogeneous material
properties of the aluminum alloy. As the feed rate can only
determine the overlap of the tool prole in radial direction,
it does not contribute to the amplitude. Therefore, the
actual relative tool-work vibration can be regarded as constant with increasing feed rate for each material under an
identical spindle speed.
Fig. 4 presents the relationship between the equivalent
amplitude and depth of cut under feed rate of 40 mm/min.
It is known that depth of cut determines the contact condition between tool and workpiece and it would have a
direct effect on cutting force resulting in the relative vibration. However, the depth of cut in SPDT is too small to
inuence the cutting force and the friction condition
between tool and chip. So no direct relationship between
depth of cut and the relative vibration was found according
to the results shown in Fig. 4, which indicates that the

equivalent amplitude is maintained within a smaller range


as the depth of cut is changed. In other words, the relative
vibration during cutting process is almost unaffected by
the change in depth of cut for each material.
During cutting process, the increase of spindle speed
can augment spindle vibration contributing to the relative
vibration between tool and workpiece, but also increases
the cutting times at the same location to reduce the cutting
force and further reduce the relative vibration. Fig. 5 shows
the effect of spindle speed on the relative vibration at different feed rates and depths of cut for three materials.
Fig. 5(a) shows that the equivalent amplitude has no distinct change under a certain spindle speed for NiP, but
changes from less than 5 nm to near 15 nm with increasing
spindle speed, which means spindle speed affects the relative vibration more signicantly than feed rate and depth
of cut for NiP. It can be explained that spindle speed mainly
inuences spindle vibration in cutting homogeneous NiP.
The evaluated equivalent amplitude is found to remain in
the range of 1018 nm for machining Cu at different spindle speeds, as shown in Fig. 5(b). During the cutting of Cu
material, the increase of spindle speed may lead to larger
spindle vibration and smaller cutting force at the same
time. Whereas, for Al7075 shown in Fig. 5(c), there is a signicant change in the equivalent amplitude at different
spindle speeds while the equivalent amplitude decreases
with increasing spindle speed. It is inferred that the
increase of spindle speed mainly reduces the cutting force,
as Al7075 is a typical inhomogeneous alloy.
In general, it can be concluded that the actual relative
tool-work vibration is signicantly inuenced by the
change in spindle speed, but not by feed rate and depth
of cut. Thus, among the cutting parameters, we only need
to take into account the effect of spindle speed on the
vibration when a model is established for predicting surface roughness.
4.3. Effect of material property
The surface topography and surface prole of the
machined surfaces are shown in Fig. 6 for every material
under same cutting parameters (spindle speed: 1000 r/
min, feed rate: 25 mm/min, depth of cut: 2 lm). A dashdot-line was drawn by connecting a series of tool nose
locations in the prole to observe the relative tool-work
vibration directly. It can be observed that the dash-dot-line
is almost straight for material NiP, which implies that the
measured prole is close to the ideal prole and the relative tool-work vibration is negligible. Moreover, based on

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

40
30

NiP

20

Cu
Al7075

10
0
25

30

35

20
15
10
NiP

Cu
Al7075

40

25

30

(a)

Feed rate (mm/min)

Equivalent amplitude (nm)

50

Equivalent amplitude (nm)

Equivalent amplitude (nm)

24

35

20
15
10
NiP

Cu
Al7075

40

25

30

(b)

Feed rate (mm/min)

35

40

(c)

Feed rate (mm/min)

40
30

NiP
Cu
Al7075

20
10
0
2

Depth of cut (m)

30

Equivalent amplitude (nm)

50

Equivalent amplitude (nm)

Equivalent amplitude (nm)

Fig. 3. Relationship between the equivalent amplitude and feed rate. (a) Spindle speed: 1000 r/min; (b) spindle speed: 1500 r/min; (c) spindle speed:
2000 r/min.

NiP

25

Cu
Al7075

20
15
10
5
0

(a)

Depth of cut (m)

20
15
10
NiP

Cu
Al7075

0
2

(b)

Depth of cut (m)

(c)

Fig. 4. Relationship between equivalent amplitude and depth of cut. (a) Spindle speed: 1000 r/min; (b) spindle speed: 1500 r/min; (c) spindle speed:
2000 r/min.

20

15

10

0
1000

1200

1400

1600

1800

Spindle speed (r/min)

2000

(a)

50

Equivalent amplitude (nm)

Equivalent amplitude (nm)

Equivalent amplitude (nm)

20

15

10

0
1000

1200

1400

1600

1800

Spindle speed (r/min)

2000

(b)

40
30
20
10
0
1000

1200

1400

1600

1800

Spindle speed (r/min)

2000

(c)

Fig. 5. Relationship between equivalent amplitude and spindle speed. (a) Material: NiP; (b) material: Cu; (c) material: Al7075.

analyzed results, the equivalent amplitude was less than


10 nm for a spindle speed of 1000 r/min, as shown in
Fig. 7. This phenomenon can be explained from the material property of NiP shown in Fig. 8. NiP has amorphous
structure, in which the atomic arrangement does not feature with long-range order and translation cyclicity. As a
result, the solid solution structure of NiP is homogeneous
and does not have grain boundaries, dislocations, twin
crystals and other defects. In addition, the passivating lm
covered on the surface of NiP basal body also has homogeneous structure without any dislocation. Therefore, the
relative vibration is very small in the case of machining
NiP because of the high isotropy of this material.
It can be observed from Fig. 6(b), that there is a large
uctuation for the displacement between tool and workpiece when machining Cu as compared to NiP. The results
of equivalent amplitude also support this conclusion and it
can be seen that, the magnitude of all the values was
greater than 10 nm, as shown in Fig. 7. This fact can be

explained on the basis that copper (Cu) is a polycrystalline


alloy material and does not have a similar homogeneous
structure as NiP. However, copper has some deoxidizing
elements and other elements in the bulk material besides
pure Cu in order to improve material performance, as seen
from the metallurgical structure of Cu shown in Fig. 9.
As shown in Table 2, Al, Zn, Mg and Cu are the main elements that make up Al7075. It can form solid solution
structure or a chemical compound in the basal body of Al
termed as trapped phase, which has a different physical
and mechanical property from the basal body. As a result,
this material was strengthened by the trapped phase,
thereby increasing the hardness. However, the existence
of trapped phase makes the material non-homogeneous.
The metallurgical structure in Fig. 9 shows the trapped
phase in Al7075 has a higher proportion and a bigger size
as compared to Cu. The machined surface prole of
Al7075 exhibits the largest displacement between tool
and workpiece along with the largest equivalent

25

Micrometers

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

(a)

Millimeters

(b)

Millimeters

(c)

Micrometers

Micrometers

Millimeters

Fig. 6. Three-dimensional topography and surface prole of machined surface. (a) Material: NiP; (b) material: Cu; (c) material: Al7075.

50

Equivalent amplitude (nm)

Intensity (cps)
2000

40

30

1500

20
1000
10
500
0

NiP
Fig. 7. Relationship
property.

Cu

Al7075

between equivalent amplitude and

material

0
20

40

60

80

100

2theta (deg.)

amplitude, i.e., Al7075 exhibits the biggest relative toolwork vibration among the three materials as shown in
Figs. 6 (c) and 7. It can be concluded that material properties can largely inuence the actual relative tool-work
vibration in SPDT, thus we must consider the effect of
material properties on the vibration behavior when predicting surface roughness.

Fig. 8. X-ray spectra of NiP.

Trapped phase

Trapped phase

4.4. Effect of tool wear


In the machining process for Al7075, serious wear of
diamond tool occurred due to its higher hardness. Fig. 10
shows the degree of tool wear when the machining was
completed. It can be seen that the circular part of diamond

Basal body

Basal body

Fig. 9. Metallurgical structure of Cu (left) and Al7075 (right).

26

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

Table 2
Chemical composition of Al7075.
Composition

Cu

Si

Fe

Mn

Mg

Zn

Cr

Ti

Al

Mass percent (%)

1.22.0

0.4

0.5

0.3

2.12.9

5.16.1

0.4

0.06

90

tool contacting the workpiece during the machining has


worn out into a straight shape. Further analysis of Fig. 10
revealed that the worn part had a height of 7.6 lm and a
width of 95.8 lm. It is generally agreed that tool wear
may lead to a larger relative tool-work vibration. On the
contrary, it is found in this study that the equivalent
amplitude becomes smaller with the increase in cutting
distance and with the aggravation in tool wear, as shown
in Fig. 11. The equivalent amplitude decreases from
44 nm to 4 nm, indicating that the relative tool-work
vibration was reduced gradually.
This can be explained from the metallurgical structure
shown in Fig. 9 (right), in which the size of trapped phase
is about 20 lm in material of Al7075, while the length of
contact zone between tool and workpiece during the turning process is not larger than 10 lm. Thus, the diamond
tool has to cut the hard trapped phase and soft basal body
of Al alternately, resulting in the repeating change of cutting state and cutting force in the turning process. The relative tool-work vibration for cutting this material must be
greater than cutting homogeneous materials. However,
when the diamond tool wears out, the length of contact
zone may increase to 100 lm, as shown in Fig. 10.
Accordingly, the part of workpiece in contact with the diamond tool is composed of the hard trapped phase and soft
basal body simultaneously, leading to a steadier material
removal process as compared to before the tool wear.
Therefore, the relative tool-work vibration gets smaller
with increase in tool wear, and the effect of tool wear on
the vibration must be considered in the turning process.

the swelling increases the peaks height of the feed components in spectrum. However, in some cases, the swelling
effect was found to decrease the surface roughness when
plastic ow for ductile materials is overwhelmed by the
effect of materials recovery [3]. The amount of recovery
is decided by the material properties and forces on the
ank face [13]. Previous research implies that the amount
of swelling depends upon the properties of the material
being cut. Softer and more ductile material show higher
swelling of the tool marks [9]. In order to quantify the
swelling effect, Sata et al. [9] dened the swelling ratio
as the ratio of power between the rst order feed component of the measured roughness spectrum and the ideal
roughness spectrum. Then, Cheung et al. [13] proposed a
local swelling ratio SRi at the ith radial section of the
machined surface which is dened as the square root of
the ratio of the power spectral density for the rst feed
components of the measured and the ideal surface roughness spectrum.
In present study, the effect of material swelling on surface roughness is required to consider in the prediction
model, so it is not feasible to process the measured roughness prole by means of spectrum method. Therefore, a
swelling proportion SP was proposed to quantify the swelling effect based directly on measured roughness prole. It
is dened as a proportion between the average height of
tool mark on measured surface and the height of ideal tool
mark:

5. The material swelling effect

where Hri is the height of the ith tool mark after both recovering and plastic owing on machined surface, as shown
in Fig. 12, Hc is the calculated height of ideal tool mark
Hc  s2 =8R; s is tool feed per work revolution, R is the tool
nose radius) and n is the number of tool marks evaluated. It
can be deduced that the effect of plastic ow will be larger
than the effect of recovery on roughness prole and thus to
increase the surface roughness, if SP > 1; while the effect of
plastic ow will be smaller than the effect of recovery on
roughness prole and thus to reduce the surface roughness, if SP < 1.

5.1. Method of quantifying the swelling effect


In the SPDT process, as the tool has two edges including
cutting edge and burnishing edge, the latter burnishes and
indents the freshly machined surface besides cutting the
materials via cutting edge in turning process. Meanwhile,
the metal left behind the cutting edge undergoes high
pressure, which results in a material ow toward the side
of the active cutting edge [16]. The material left behind
the ank face recovers after burnishing [3]. Moreover, the
cutting force along the main cutting edge pushes aside
the work material near the tool nose, causing it to ow
toward the free surface [17]. The combined effect of plastic
ow, burnishing, and elastic recovery is called the swelling
effect [13], which causes the change of tool marks generation on the machined surface, as shown in Fig. 12.
Liu et al. [18] showed plastic side ow could increase
peak-to-valley roughness due to the material piling up at
the trailing edge of the tool. Sata et al. [9] found work
material swells at the end of the active cutting edge causing a greater tool mark on the machined surface because

Pn
SP

i1 Hri
nHc

5.2. Effect of cutting parameters


Sata et al. [9] studied the swelling ratio of C45 and brass
at the different feed rates. It was found that the swelling
ratios remain nearly constant within a certain feed rate
range. Cheung et al. [13] analyzed the distribution of
swelling ratio on machined surface of aluminum single
crystal and Al6061 to evaluate the materials anisotropy.
However, no report was found to predict surface roughness
taking into account the swelling effect. Therefore, the
effect of cutting parameters on the material swelling and

27

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

Fig. 10. Image of tool wear (left, magnication 200) and size of worn part (right, magnication 1000).

1.5

40

Swelling proportion

Equivalent amplitude (nm)

50

30

20

10

0
5.5

6.0

6.5

7.0

7.5

8.0

8.5

0.5
1000

Cutting distance (km)

1200

1400

1600

1800

2000

Spindle speed (r/min)


Fig. 11. Relationship between equivalent amplitude and tool wear.
Fig. 13. The swelling ratio of NiP at different cutting parameters.

Hc

Hf

Hr

Ideal surface calculated


Surface after plastic flow
Surface after recovery

Fig. 12. Schematic illustration for the effect of swelling on surface


generation (Hc height of ideal tool mark calculated, Hf height of tool
mark after plastic ow, Hr height of tool mark after recovery).

the change of surface roughness caused by the swelling


effect is still far from understand very well and needs to
be further studied.
For the purpose of correlating a prediction model of surface roughness with the swelling effect, it was analyzed at
different cutting parameters for NiP, as shown in Fig. 13.
The swelling proportion is in the ranges of 1.211.39,
1.041.11 and 0.961.08 at varying spindle speeds of
1000 r/min, 1500 r/min and 2000 r/min respectively under
different feed rates and depths of cut. The result implies
that the swelling proportion nearly keeps invariable with
the changing feed rates and cutting depths for a xed spindle speed. However, it decreases obviously when the spindle speed was increased, which means machined surface

becomes smoother at a higher spindle speed. This can be


explained based on the fact that at higher spindle speed
the same position on the fresh machined surface will be
burnished and indented many times, which cause a larger
residual stress onto the surface, with a larger recovery and
thus a lower height of tool mark. The results reveal that the
materials swelling effect was mainly affected by the
spindle speed, which is very useful and signicant for
the proposed prediction model in this study, because the
roughness prole after plastic ow and recovery can be
clearly known based on the swelling proportion at different spindle speeds.
6. A prediction model of surface roughness
6.1. Creation of a prediction model
Based on above analysis, the main factors affecting the
relative tool-work vibration are spindle speed, material
property and tool wear. In addition, the swelling effect of
a material will change mainly with the change in spindle
speed. Therefore, in the prediction model, experiments
are necessary to determine the relative tool-work vibration
and the swelling effect at different spindle speeds for each
material. As shown in Fig. 14, the roughness prole was
rst measured on the machined surface in radial direction,
and then processed by the methods discussed in the

28

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

Experiments

Equivalent
vibration

Cutting
parameters

Data
processing

Materials
property
Relative toolwork vibration

Then, both the roughness prole after the swelling


effect Y S x and the curve of equivalent simple harmonic
motion Y h x were produced with the method of addition
of waveforms on MATLAB/Simulink software, where Y h x
is given by Eq. (4) and shown in Fig. 15(c). So, the addition
of two waveforms resulted in a new roughness prole Y(x),
as shown in Fig. 15(d). Finally, predicted arithmetic roughness value Ra was calculated with export data of the roughness prole Y(x) on MATLAB/Simulink software. It is noted
that the roughness prole Y(x) contains both the relative
tool-work vibration and the swelling effect, is only used
to calculate roughness value but not actual roughness prole of machined surface.
In conclusion, there are three aspects of signicant feature in the prediction model presented in this paper, as
follows:

Measure roughness profile

The swelling
proportion

Ideal roughness
profile

Roughness profile after swelling


effect

Addition of waveforms and data processing


Surface roughness

(1) The evaluated vibration is more credible and close to


the actual vibration in the turning process because
of the relative tool-work vibration coming from the
machined surface.
(2) The swelling effect, i.e., the material plastic ow and
recovery in turning is rst taken into account in the
prediction of surface roughness.
(3) This prediction model adopts the approach combining
machining
theory
with
experimental
investigation.

Fig. 14. A block diagram of the prediction model of surface roughness.

previous section. As a result, the actual relative tool-work


vibration was simplied as a simple harmonic motion to
 while swelling procalculate the equivalent amplitude A,
portion SP of each material can be determined according
to Eq. (9).
Once the cutting parameters are conrmed, the ideal
roughness prole in radial direction shown in Fig. 15(a)
can be expressed as

6.2. Verication of the prediction model

Y I x

x2
2R

Y I x

x  ns2
2R

if 0 6 x 6 s=2

10

if x > s=2

Verication tests of face cutting were carried out for


materials NiP and Cu. The tests were conducted on a
four-axis CNC ultra-precision machine tool (made by
Nachi-Fujikoshi Corp., Japan) and a diamond tool used in
tests has a rake angle of 0, a front clearance angle of 6
and a tool-nose radius of 0.5 mm. The cutting conditions
are tabulated as Table 3. By using the model proposed in

11

where s is tool feed per work revolution, R is the tool nose


radius and n = dx s=2=sede means rounding down to
the nearest whole unit). Since the swelling effect cannot
be avoided, both plastic ow and recovery of the material
were taken into account to evaluate a roughness prole
after swelling effect Y S x, which is calculated on the basis
of the ideal roughness prole as well as the calculated
swelling proportion SP, as shown in Fig. 15 (b).

Y S x SPY I x

the present study, the equivalent amplitude A and swelling


proportion SP were rst evaluated under varying cutting
parameters for NiP and Cu, as shown in Table 4. Then surface roughness values Ra were predicted.
The predicted values, measured values and ideal values
are shown in Fig. 16, in which the ideal values were calculated from the ideal roughness prole Y I s given by Eqs.
(10) and (11). It can be seen that there are three kinds of

12

Y1 (x)

Y(x)

(a)

(d)

Hc
0

2s

3s

Addition of waveforms

After swelling effect

Ys(x)

Yh(x)

(b)

Hr
Hc
0

(c)

2A

2s

3s

Fig. 15. The schematic diagram of predicting process of surface roughness. (a) Ideal roughness prole; (b) roughness prole after swelling effect, Hc and Hr
shown in Fig. 12; (c) equivalent simple harmonic motion; (d) roughness prole containing both the relative tool-work vibration and the swelling effect.

29

J. Chen, Q. Zhao / Measurement 69 (2015) 2030


Table 3
Cutting conditions of verication tests.
Term
no.

Spindle speed
(rpm)

Feed rate
(mm/min)

Depth of cut
(lm)

Tool-nose
radius (mm)

Term
no.

Spindle speed
(rpm)

Feed rate
(mm/min)

Depth of cut
(lm)

Tool-nose
radius (mm)

1
2
3
4
5
6
7

1500
1500
1500
1500
1500
1500
1500

25
30
35
40
40
40
40

2
2
2
2
4
6
8

0.5
0.5
0.5
0.5
0.5
0.5
0.5

8
9
10
11
12
13
14

2000
2000
2000
2000
2000
2000
2000

25
30
35
40
40
40
40

2
2
2
2
4
6
8

0.5
0.5
0.5
0.5
0.5
0.5
0.5

Table 4
Equivalent amplitude A and swelling proportion SP calculated by the
model.
A (nm) and SP

Spindle speed (rpm)


1500

2000

12

1.11

1.08

Material NiP
A
SP
Material Cu
A
SP

13

14

0.82

0.85

surface roughness values, which increases with the


increase in feed rate, as illustrated in Fig. 16(a) and (c).
This can be explained that a bigger feed rate will make
the tool marks deeper and wider in ideal roughness prole,

and lead to a larger ideal value of surface roughness as


compared to a smaller feed rate. Additionally, it was found
that there was a good accordance between the predicted
and the measured values, indicating the prediction error
was small or close to negligible in this case.
Fig. 16(b) and (d) presents the measured surface roughness having no obvious uctuation when the depth of cut
increases, which supports that the effect of depth of cut
on the relative vibration and the swelling effect can be
ignored in the prediction model. The little uctuation for
measured surface roughness may be caused due to the
change of the machine tool error or the swelling proportion with the change in depth of cut. However, the results
for the measured and predicted values were in good
agreement.
It is noticed that the two materials (NiP and Cu) were
machined with the same diamond tool, the same machine
tool and the same cutting parameters, theoretically,

70

60

Surface roughness R a (nm)

Surface roughness R a (nm)

70

50
40
30

NiP
Cu

20
10
0
25

Ideal
35

40
30
20

NiP
Cu
Ideal

10

40

(a)

(b)

Depth of cut (m)

40

40

Surface roughness R a (nm)

Surface roughness R a (nm)

50

0
30

Feed rate (mm/min)

30

20

NiP
Cu
Ideal

10

0
25

60

30

35

Feed rate (mm/min)

40

(c)

30

20

NiP
Cu
Ideal

10

Depth of cut (m)

(d)

Fig. 16. Model predicted surface roughness (dash dot line), measured surface roughness (solid line) and ideal surface roughness for NiP and Cu. (a) Spindle
speed: 1500 r/min, depth of cut: 2 lm; (b) spindle speed: 1500 r/min, feed rate: 40 mm/min; (c) spindle speed: 2000 r/min, depth of cut: 2 lm; (d) spindle
speed: 2000 r/min, feed rate: 40 mm/min.

30

J. Chen, Q. Zhao / Measurement 69 (2015) 2030

the ideal surface should be identical, but the measured surface roughness values were found to have a big difference
in comparison to the corresponded ideal value. This difference could be caused by the effect of material property on
the relative vibration and the swelling effect. Especially for
material Cu, most of predicted surface roughness and measured surface roughness are less than ideal value, which is
caused by the swelling proportion SP < 1, i.e., Hr < Hc in
Fig. 15(b). Therefore, it is necessary to consider the inuence of material property in order to achieve high prediction accuracy when predicting the surface roughness.
Overall, using the prediction model in the present study,
a good accordance between the measured and predicted
values is realized. The average prediction error of surface
roughness Ra is found to be 5.1% as well as the error is
within 6.5% in most cases.
7. Conclusions
The actual relative tool-work vibration during the turning process is different from the relative vibration measured before turning. The swelling effect can obviously
affect the surface roughness prole through changing the
size of tool mark, thus have to be taken into account the
prediction of surface roughness in SPDT.
In machining a material with a homogeneous structure
(e.g., NiP), the actual relative tool-work vibration caused
was very smaller when compared to inhomogeneous
materials. This is due to the fact that homogeneous materials do not present grain boundaries, dislocations, compound twins and other defects. But for inhomogeneous
materials (e.g., Cu and Al7075), material induced vibration
was larger and mainly determined by the size of trapped
phase and the length of contact zone between tool and
workpiece.
Among three types of cutting parameters, the spindle
speed has the most dominant inuence on the relative
vibration and the swelling effect when compared to that
of the other two parameters of feed rate and depth of cut.
Moreover, it was found that a higher spindle speed leads
to a lower swelling proportion and smoother machined surface, because the larger residual stress and amount of
recovery can be caused on the fresh machined surface.
Using the approach combining machining theory with
experimental investigation, a prediction model of surface
roughness in SPDT was proposed. It takes into account
the actual relative tool-work vibration extracted from the
machined surface and the swelling effect, which represents
the complicated elastic and plastic deformation in cutting.
Therefore, this model is further close to the actual cutting
process.
There is a good agreement between the model predicted
and the measured values. The average prediction error of
surface roughness Ra is found to be 5.1% while in most cases

the error is within 6.5%, which further proves the improvement of the model proposed in the present study.
Acknowledgments
The authors would like to express their sincere thanks
to the National Scientic Foundation of China (NSFC)
(Contract No. 51205343) and Postdoctoral Science
Foundation of China (Contract No. 2012M520595) for their
nancial support of the research work.
References
[1] M. Tauhiduzzaman, S.C. Veldhuis, Effect of material microstructure
and tool geometry on surface generation in single point diamond
turning, Precis. Eng. 38 (2014) 481491.
[2] P.G. Benardos, G.C. Vosniakos, Predicting surface roughness in
machining: a review, Int. J. Mach. Tools Manuf. 43 (2003) 833844.
[3] M.C. Kong, W.B. Lee, C.F. Cheung, S. To, A study of materials swelling
and recovery in single-point diamond turning of ductile materials, J.
Mater. Process. Technol. 180 (2006) 210215.
[4] K.V. Rao, B.S.N. Murthy, N. Mohan Rao, Prediction of cutting tool
wear, surface roughness and vibration of work piece in boring of AISI
316 steel with articial neural network, Measurement 51 (2014) 63
70.
[5] Z. Hessainia, A. Belbah, M.A. Yallese, T. Mabrouki, J.F. Rigal, On the
prediction of surface roughness in the hard turning based on cutting
parameters and tool vibrations, Measurement 46 (2013) 16711681.
[6] V. Upadhyay, P.K. Jain, N.K. Mehta, In-process prediction of surface
roughness in turning of Ti6Al4V alloy using cutting parameters
and vibration signals, Measurement 46 (2013) 154160.
[7] A.M. Zain, H. Haron, S. Sharif, Prediction of surface roughness in the
end milling machining using articial neural network, Expert Syst.
Appl. 37 (2010) 17551768.
[8] D. Karayel, Prediction and control of surface roughness in CNC lathe
using articial neural network, J. Mater. Process. Technol. 209 (2009)
31253137.
[9] T. Sata, M. Li, S. Takata, H. Hiraoka, Q.Q. Li, X.Z. Xing, X.G. Xiao,
Analysis of surface roughness generation in turning operation and its
applications, Annals CIRP 34 (1985) 473476.
[10] S. Takasu, M. Masuda, T. Nishiguchi, Inuence of study vibration
with small amplitude upon surface roughness in diamond
machining, Annals CIRP 34 (1985) 463467.
[11] C.F. Cheung, W.B. Lee, Modelling and simulation of surface
topography in ultra-precision diamond turning, Proc. Inst. Mech.
Eng. Part B: J. Eng. Manuf. 214 (2000) 463480.
[12] W.B. Lee, C.F. Cheung, A dynamic surface topography model for the
prediction of nano-surface generation in ultra-precision machining,
Int. J. Mech. Sci. 43 (2001) 961991.
[13] C.F. Cheung, W.B. Lee, A multi-spectrum analysis of surface
roughness formation in ultra-precision machining, Precis. Eng. 24
(2000) 7787.
[14] H. Wang, S. To, C.Y. Chan, C.F. Cheung, W.B. Lee, A theoretical and
experimental investigation of the tool-tip vibration and its inuence
upon surface generation in single-point diamond turning, Int. J.
Mach. Tools Manuf. 50 (2001) 241252.
[15] W.B. Lee, C.F. Cheung, S. To, Materials induced vibration in ultraprecision machining, J. Mater. Process. Technol. 8990 (1999) 318
325.
[16] M.C. Shaw, J.A. Crowell, Finish machining, Annals CIRP 13 (1965)
(1965) 522.
[17] T. Sata, Surface nish in metal cutting, Annals CIRP 12 (1964) 190
197.
[18] K. Liu, S.N. Melkote, Effect of plastic side ow on surface roughness
in micro-turning process, Int. J. Mach. Tools Manuf. 46 (2006) 1778
1785.

Вам также может понравиться