You are on page 1of 8

Precision Engineering 41 (2015) 145152

Contents lists available at ScienceDirect

Precision Engineering
journal homepage: www.elsevier.com/locate/precision

Technical note

Nanometric cutting in a scanning electron microscope


Fengzhou Fang a,b , Bing Liu a,b , Zongwei Xu a,
a
b

State Key Laboratory of Precision Measuring Technology & Instruments, Tianjin University, Tianjin 300072, China
School of Mechanical Engineering, Tianjin University, Tianjin 300072, China

a r t i c l e

i n f o

Article history:
Received 25 September 2014
Received in revised form 29 January 2015
Accepted 30 January 2015
Available online 12 February 2015
Keywords:
Nanometric cutting
Device
Scanning electron microscope
Online observation
Tool edge radius
Focused ion beam

a b s t r a c t
A nanometric cutting device under high vacuum conditions in a scanning electron microscope (SEM)
was developed. The performance, tool-sample positioning, and processing capacity of the nanometric
cutting platform were studied. The proposed device can be used to realize a displacement of 7 m, with
a closed-loop resolution of 0.6 nm in both the cutting direction and the depth direction. Using a diamond
cutting tool with an edge radius of 43 nm formed by focused ion beam (FIB) processing, nanometric
cutting experiments on monocrystalline silicon were performed on the developed cutting device under
SEM online observation. Chips and machining results of different depths of cut were studied during
the cutting process, and cutting depths of less than 10 nm could be obtained with high repeatability.
Moreover, the cutting speed was found to exhibit a strong relationship with the brittleductile transition
depth on brittle material. The experimental results of taper cutting and sinusoidal cutting indicated that
the developed device has the ability to perform multiple degrees of freedom (DOFs) cutting and to study
nanoscale material removal behaviour.
2015 Elsevier Inc. All rights reserved.

1. Introduction
Along with the development of cutting technology from conventional cutting to micro-cutting, even to nanometric cutting, the
depth of cut is decreasing and machining accuracy is improving.
Due to the size effect, the material removal mechanism of nanometric cutting may be different from that of conventional cutting.
Yuan et al. [1] studied the effects of the tool edge radius on the
minimum cutting thickness via diamond cutting of Al-alloys, indicating that when the edge radius of the diamond cutting tool was
0.20.6 m, the minimum thickness of the chips was observed to
be 0.050.2 m. Fang et al. [25] studied the nanometric cutting
mechanism of monocrystalline silicon and found that the chip formation in nanometric machining is based on extrusion rather than
shearing. This result indicates that in nanoscale machining, the cutting mechanism is signicantly different from that in conventional
cutting. Malekiana [6] found the minimum cutting thickness was a
function of the tool edge radius and the friction coefcient; depending on the tool geometry and the workpiece property, the average
minimum cutting thickness was 0.23 times the tool edge radius.
Wu et al. [7] resorted to scratch experiments to study the inuence of plastic machining of monocrystalline silicon with different
crystal orientations, and the phase transformation was analyzed

Corresponding author. Tel.: +86 22 27403753 3.


E-mail address: zongweixu@tju.edu.cn (Z. Xu).
http://dx.doi.org/10.1016/j.precisioneng.2015.01.009
0141-6359/ 2015 Elsevier Inc. All rights reserved.

using Raman spectroscopy. Many research results indicated that


the conventional cutting theory has been unable to interpret the
results and phenomena effectively during the nanometric cutting
process.
In recent years, researchers have focused their efforts on the
study of the nanometric cutting mechanism, and many remarkable achievements have been obtained. With the development
of computer technology, research studies on the nanometric cutting mechanism were widely implemented by molecular dynamics
(MD) simulation [811]. To verify the MD simulation results and
experimentally clarify the nanometric cutting mechanism, the
atomic force microscopy (AFM) based nanoscratching method
[1215] and the ultra-precision turning based cutting experiments
[1618] were considered.
However, in the study of the nano-cutting mechanism based
on AFM or nanoscratching, although the AFM probe can be quite
sharp, the probe or scratch head structure parameters were different from the tool parameters, such as the difference in the rake
face and ank face, thereby reducing the validity and authenticity of the examined nano-cutting mechanism. Moreover, during
ultra-precision turning-based cutting experiments, the depth of
cut was difcult to achieve on the nanoscale and it was difcult
to detect when the nanoscale cutting chip comes out. Controlling
the cutting depth to nanometer precision is essential but not sufcient. To fully understand the governing physics of nanometric
cutting, the cutting process and chip morphology must be observed
online.

146

F. Fang et al. / Precision Engineering 41 (2015) 145152

Fig. 1. Schematic of the nanometric cutting device enabling online SEM observation.

The key task in the study of nanometric cutting mechanism is


to develop a device for high resolution in situ and online observation of nanometric cutting. The SEM technique has undergone
considerable development in recent years because of its high resolution, with many experiments involving the characterization of
the properties of nanoscale materials, in situ indentation experiments [1922], and nano-manipulation experiments [23,24]. FIB
technology has also been widely used in the nano-manufacturing
eld, especially in the fabrication of diamond tools with a nanoedge radius and versatile edge prole [25,26]. Therefore, in this
paper, a nanometric cutting device inside an SEM was developed
and tested to study the nanometric cutting mechanism. Nanocutting experiments were performed to evaluate the performance
of the device.
2. Nanometric cutting device
As shown in Fig. 1, the nanometric cutting device is attached to
the sample stage of an SEM, which can be disassembled for testing
and maintenance. The device consists of multi-DOFs micropositioners, a multi-DOFs nanoscale motion stage, and a nano-manipulator.
In addition, the sensors for tool setting, diamond cutting tool, tool
handle and other adapting workpieces are included in the device.
The main parts of the device are made of stainless steel, which
meets the requirement of being stiff and nonmagnetic.
One of the most important requirements for in situ observation
of the nanometric cutting process is high-resolution SEM imaging.
In all experimental processes, both the coarse and ne positioning are observed via SEM imaging with nanometer resolution. The

environment of the sample chamber is a vacuum with a vibration


isolated platform that is capable of preventing the effects of foreign substances and vibration. The function of beam shift in the
SEM enables the offset range of 60 m via electron deection of
the SEM image.
The micropositioner system makes use of the FIB stage, which
belongs to the FIB dual beam system used in this study. The system consists of two parts: micropositioner A and micropositioner B.
For micropositioner A, there are three DOFs (XYT-axis). For micropositioner B, there are another two DOFs (ZR-axis) besides A. The
sample holder is xed on the micropositioner B, and then, the sample is glued onto the sample holder using liquid carbon conductive
adhesive, avoiding the inuence of sticky glue on the stiffness of
the cutting process. Micropositioner A has a range of movement of
approximately 20 mm to 20 mm, which is capable of taking the
sample rapidly to the machining area, where the gap between the
diamond cutting tool and the sample is less than 7 m. The Z-axis is
parallel to the SEM electron beam, which is used to adjust the relative height between the cutting tool and the sample. The R-axis can
drive the sample to rotate at a certain angle, which is used for angle
adjustment between the cutting direction and sample surface. The
rotation T-axis can be used to realize the online observation of different angles of the nanometric cutting process via the use of the
SEM. Overall, the task of the micropositioner is coarse-positioning
the sample and adjusting the view direction.
A nanoscale motion stage of three DOFs (XYZ-axis), with dimensions of 30 mm 30 mm 42 mm, is mounted on the isolation
platform for the task of ne-positioning the cutting tool and generating the crossfeed and infeed cutting motions. Each axis of this
stage makes use of a piezoelectric ceramic (PZT) actuator, which
drives the tool to accomplish a maximum cutting depth of 7 m
with a closed-loop resolution of 0.6 nm and length of cut of 7 m.
Both the crossfeed stiffness and infeed stiffness of the stage are
specied more than 6 N/m. The stiffness of the motion stage in
the infeed direction is more important than that in the crossfeed
direction. The total stiffness of the instrument was analyzed according to nite element analysis, and it is about 2 N/m in the infeed
direction. The nanoscale motion stage positions the diamond cutting tool against the sample exactly under SEM online observation.
The diamond tool can be positioned in the XYZ DOFs.
After each cutting experiment, a fraction of the chips remained
on the rake face of the tool, affecting the subsequent observation
of the material removal process. For better observation, a MM3A
nano-manipulator is used to remove the remaining chips from
the rake face. Each axis of the nano-manipulator makes use of
an piezoelectric actuator, enabling a step size of 0.5 nm. The positioning accuracies presented in Table 1 proved to be sufcient for
the tasks mentioned. The positioning accuracies are specied as
follows.
3. Performance test
In this study, the nanoscale motion stage driven by a PZT
actuator is integrated into the SEM. Due to the creep properties
of PZT, a closed-loop control system is used to eliminate creep

Table 1
Positioning accuracies during different tasks of the nanometric cutting experiments (manufacturers specications).
Task

Actuator

Travel range

Coarse positioning of sample


Fine positioning of cutting tool
Removal of the chips
Positioning the cutting tool to the SEM visual eld
Angle adjustment of the sample

Micropositioner
Nanoscalemotion stage
Nano-manipulator
Beam shift of SEM
Stage platform

25 mm
7 m
120 ; 12 mm
60 m
360

Positioning accuracy
X

20 m
3 nm

20 m
3 nm

20 m
3 nm
0.25 nm

10e7 rad
0.1

F. Fang et al. / Precision Engineering 41 (2015) 145152

147

Table 2
Comparison of the command and the actual depth of cut. (Average of three repeated
measurements).
Command depth of
cut (nm)

Actual depth of
cut (nm)

Maximum error
value (nm)

10
50
100

10.6 2
54.9 2
105.4 2

2.6
6.9
7.4

displacement. However, when the depth of cut is on the nano-scale,


the most important issues for the device are positioning accuracy
and stability. In consideration of the signicant inuence on the
cutting results, the stability of the nanoscale motion stage was
analyzed to ensure that the actual depth of cut is the same as the
command depth of cut.
First, the device developed was placed under working conditions for more than 3 h to ensure that the device was stable. Second,
a single crystal diamond tool with straight edge was used to cut a
single crystal copper sample, which was processed into a step structure of different depths, as shown in Fig. 2. Single crystal copper was
used as the sample rather than silicon because the large depth of
cut may cause the silicon crack, thereby affecting the measurement
results. Finally, the height difference between the two machined
surfaces (1) and (2) was measured using a White Light Interferometer. Fig. 3 shows the measurement results of the height difference,
i.e., the actual depth of cut.
From the results above, it is known that when the command
depth of cut was 10 nm, 50 nm and 100 nm, the measured actual
depth of cut was 10.6 nm, 54.9 nm and 105.4 nm, respectively.
The comparison of the command depth of cut and actual depth
is listed in Table 2. The difference between the command depth
of cut and the measured depth of cut may result from different sources, including the hysteresis of the piezoelectric ceramics,
electric noise, measurement uncertainties, etc. The observed differences can be used for compensating the error in the actual depth
of cut.

4. Nano-cutting experiments under real-time SEM


observation
The nano-cutting experiments were performed in a
temperature-controlled metrology laboratory. To better understand the cutting process, the fabrication method of the diamond
cutting tool and sample must be known.

Fig. 3. Measurement results of the height difference between adjacent steps


(machined surfaces (1) and (2)). The hills in between the surfaces (1) and (2) come
from the surface of the chip adhering to the workpiece.

Fig. 4. Diamond cutting tool shaped by FIB processing.


Fig. 2. Step structure of the single crystal copper sample.

148

F. Fang et al. / Precision Engineering 41 (2015) 145152

Fig. 5. SEM measurement of the tool edge radius.

4.1. Diamond cutting tool


A diamond cutting tool with a straight edge was used to perform
the nanometric cutting experiments. The device was integrated
into a FIB dual beam system. Firstly, the diamond cutting tool was
fabricated using FIB processing, and then, the tool was used to
perform a series of experiments under SEM online observation to
study the nanometric cutting mechanism. For better observation
of the material removal process, the cutting edge is machined to
be straight with the length of 10 m using FIB processing. In the
machining process, the accelerating voltage was 30 kV and the ion
beam current was 100 pA. The diamond cutting tool as fabricated

Fig. 6. Rectangular groove structure of the monocrystalline silicon sample.

is shown in Fig. 4. The tool edge radius is approximately 43 nm, as


measured from the SEM image shown in Fig. 5. The rake angle of
the tool is 0 , and the relief angle is 8 .
4.2. Sample fabrication method
In this study, monocrystalline silicon was used as the sample
in all of the experiments described below. To avoid plough interference between the sample and the tool side surface during the
cutting process, the silicon was machined into a rectangular groove
structure to observe the material removal process clearly, as shown
in Fig. 6. The width of the rectangular groove structure is slightly

Fig. 7. Generation process of the cutting tool shadows on sample surface.

F. Fang et al. / Precision Engineering 41 (2015) 145152

149

Fig. 8. Schematic diagram of the cutting depth error.

less than the straight tool edge of 10 m and the length is more
than the range of nanoscale motion stage of 7 m.
4.3. Tool and sample contact

Fig. 10. SEM photograph of the cutting process. The cutting speed was 23.5 nm/s.

In the nano-cutting experiments, it is important to establish the


contact reference between the tool and the sample. To conrm
the tool-sample contact, the micropositioner knob was manually
rotated to move the sample towards the cutting tool until the gap
between the sample and the tool was within less than 3 microns
using the micropositioner. Next, the sample was maintained static
and the cutting tool was controlled to approach the sample slowly
by using the nanoscale motion stage with visual feedback via SEM
imaging. When the tool was close enough to the sample surface, a
shadow of the tool would appear on the sample surface. Fig. 7(ad)
shows the generation process of the shadows. Thus, it is very useful
to judge whether the tool is in contact with the sample surface.
Note that if the sample surface is not strictly parallel to the
cutting direction or cutting edge, then the cutting depth of the
machined surface is not consistent. This inconsistency would affect
the analysis accuracy of the nano-cutting results, such as Raman
measurements, and so on. The differential depth of cut between
the two ends of the machined surface is shown in Fig. 8. In the case
that the straight edge is 10 m and the cutting distance is 7 m,
the subsequent error of cutting depth d1 and d2 is 17.5 nm and
12.2 nm, respectively, assuming both and are controlled well
to within 0.1 along the T-axis and R-axis, respectively. When the
depth error and the depth of cut are of the same order of magnitude, a pre-cut procedure was used to ensure the cutting depth of
the machined surface is consistent.
In addition, it is important to ensure that both the cutting tool
and the convex structure are at the same height along the Z direction, using the method schematically shown in Fig. 9. First, the
T-axis was rotated to make the sample and the tool to tilt at a certain
angle. Next, the cutting tool was moved to the sample surface along
the Y direction, producing a tool mark. Subsequently, the distance

was measured via examination of the cross section. Therefore, the


height difference between the cutting tool and the sample can be
determined, and then, the nanoscale motion stage can be controlled
to make the tool rise or fall a displacement of Z.

Fig. 9. Schematic of the tool positioning in the Z direction.

4.4. Nano-cutting experiments


For a cutting experiment, the cutting tool was rst brought
in contact with the sample using the technique illustrated above.
Once the tool-sample contact was established, the tool was moved
along the reverse cutting direction to the edge of the sample. Next,
attempts were made at cutting the sample with a feed of 10 nm
every time. This was performed until the sample was cut, and the
location at this moment was recorded as the cutting depth of zero.
Subsequently, the cutting experiments were started by moving the
diamond tool. Figs. 10 and 11 show the cutting process viewed by
online SEM observations. Fig. 12 shows SEM micrographs of the
chips formed by using the diamond cutting tool with edge radius
of 43 nm, where the depth of cut was (a) 5 nm, (b) 15 nm, (c) 25 nm,
and (d) 40 nm. As seen in Fig. 12(ac), when the depth of cut was
smaller, the chips formed were continuous. Such continuous chips
obtained at these depths of cut indicated that the cutting was performed in ductile mode and that plastic deformation had taken
place. However, as the depth of cut increased to 40 nm (Fig. 12(d)),
discontinuous and fractured chips were obtained. This result indicates that the cutting at this depth of cut was in brittle mode.
Fig. 13 shows AFM images and SEM micrographs of nano-cutting
with an input sinusoidal trajectory (peak-to-peak value of 50 nm,

Fig. 11. SEM photograph of the entire eld of view. The undeformed cutting depth
was (a) 5 nm, (b) 10 nm, and (c) 20 nm.

150

F. Fang et al. / Precision Engineering 41 (2015) 145152

Fig. 12. SEM micrographs of chips with different depths of cut. The cutting speed was 23.5 nm/s.

Fig. 13. Cutting results with a sinusoidal trajectory half of a full period on single-crystal germanium. The cutting speed was 23.5 nm/s.

F. Fang et al. / Precision Engineering 41 (2015) 145152

Fig. 14. AFM image of a tapered cutting groove on single-crystal silicon. The cutting
speed was 23.5 nm/s.

cycle of 12 m) on single-crystal germanium. The cross-sectional


prole along the X-axis shown in Fig. 13(b and c) correspond to
the machined surface morphology. As seen in Fig. 13(c), the cutting
direction is from left to right. Both plastic cutting and brittle cutting
occurred during this nano-cutting experiment. The critical depth
of brittleductile transition for single-crystal germanium was
approximately 20 nm in this particular case (speed of 23.5 nm/s,
dry cutting, tool edge radius of 43 nm). Note that the beginning
and end of brittle area appeared at different depths. As shown in
Fig. 13(c), when the cutting tool was controlled from the sinusoidal

151

trajectorys peak to valley, compressive stress generated ahead of


the cutting tool greatly contributed to the ductile machining of
brittle material. When the cutting tool was controlled from the
sinusoidal trajectorys valley to the next peak, the tensile stress and
the adhesion between the chips and the cutting tool would degrade
the ductile machining results.
Fig. 14 shows the AFM image with an increasing depth of cut on
monocrystalline silicon. The depth of cut was from 0 to 30 nm, with
the cutting distance of 7 m. The experimental results mentioned
above indicate that multi-DOFs cutting can be realized using the
nanometric cutting device.
Fig. 15 shows SEM micrographs of the machined surface
obtained in cutting silicon, where the depth of cut was 24 nm and
the cutting speeds were (a) 1.4 mm/s, (b) 117.6 nm/s, (c) 23.5 nm/s.
The gure shows that at cutting speeds (a) and (b) the surface was
free from cracks. However, when the cutting speed was reduced to
23.5 nm/s (c), some brittle cracks appeared on the silicon surface.
This result indicates that the higher the cutting speed, the better the
surface quality. In addition, compared to the results in Fig. 12(c),
the cutting speed signicantly affected the brittleductile transition depth. Further study of the inuence of the cutting speed will
be addressed in another paper.
5. Conclusions
A new nanometric cutting device that employs a nanoscale
motion stage to provide accurate nano-cutting motions was developed. The device used inside an SEM can realize the online
observation of the nano-cutting process, which can be used
to effectively study the nanoscale material removal mechanism experimentally. The fundamental performance of the device
was investigated. A method of conrming tool-sample contact

Fig. 15. SEM micrographs of machined silicon surfaces at different cutting speeds.

152

F. Fang et al. / Precision Engineering 41 (2015) 145152

was introduced. The experimental results of nanometric cutting


demonstrated that the developed device has the ability to perform multi-DOFs cutting and enables nanoscale material removal
behaviour to be achieved.
Acknowledgments
The authors appreciate the support of the National
Basic Research Program of China (973 Program, Grant No.
2011CB706700), the National Natural Science Foundation of
China (Grant Nos. 91423101 & 51275559), LPMT, CAEP (Grant
No. KF13008), and the 111 project by the State Administration
of Foreign Experts Affairs and the Ministry of Education of China
(Grant No. B07014). The authors thank Mr. Wei Wu for support in
the experiments.
Appendix A. Supplementary data
Supplementary data associated with this article can be found,
in the online version, at http://dx.doi.org/10.1016/j.precisioneng.
2015.01.009.
References
[1] Yuan Z, Zhou M, Dong S. Effect of diamond tool sharpness on minimum cutting
thickness and cutting surface integrity in ultraprecision machining. J Mater
Proc Technol 1996;62(4):32730.
[2] Fang F, Venkatesh V. Diamond cutting of silicon with nanometric nish. CIRP
Ann: Manuf Technol 1998;47(1):459.
[3] Fang F, Wu H, Liu Y. Modelling and experimental investigation on nanometric cutting of monocrystalline silicon. Int J Mach Tools Manuf 2005;45(15):
16816.
[4] Fang F, Wu H, Zhou W, Hu XT. A study on mechanism of nano-cutting single
crystal silicon. J Mater Proc Technol 2007;184(1):40710.
[5] Fang F, Zhang G. An experimental study of edge radius effect on cutting single
crystal silicon. Int J Adv Manuf Technol 2003;22(910):7037.
[6] Malekian M, Mostofa M, Park S, Jun M. Modeling of minimum uncut chip
thickness in micro machining of aluminum. J Mater Proc Technol 2012;212(3):
5539.
[7] Wu H, Melkote SN. Effect of crystallographic orientation on ductile scribing
of crystalline silicon: role of phase transformation and slip. Mater Sci Eng A
2012;549:2005.

[8] Cheong W, Zhang L. Molecular dynamics simulation of phase transformations in silicon monocrystals due to nano-indentation. Nanotechnology
2000;11(3):173.
[9] Lai M, Zhang X, Fang F. Study on critical rake angle in nanometric cutting. Appl
Phys A 2012;108(4):80918.
[10] Pei Q, Lu C, Lee H. Large scale molecular dynamics study of nanometric machining of copper. Comput Mater Sci 2007;41(2):17785.
[11] Yang X, Guo J, Chen X, Kunieda M. Molecular dynamics simulation of the material removal mechanism in micro-EDM. Precis Eng 2011;35(1):517.
[12] Lee SH. Analysis of ductile mode and brittle transition of AFM nanomachining
of silicon. Int J Mach Tools Manuf 2012;61:719.
[13] Tong Z, Liang Y, Jiang X, Luo X. An atomistic investigation on the mechanism of
machining nanostructures when using single tip and multi-tip diamond tools.
Appl Surf Sci 2014;290:45865.
[14] Yan Y, Sun T, Dong S. Study on effects of tip geometry on AFM nanoscratching
tests. Wear 2007;262(3):47783.
[15] Zhou H, Qiu S, Zhang X, Xu C. Mechanical characteristics of soft-brittle HgCdTe
single crystals investigated using nanoindentation and nanoscratching. Appl
Surf Sci 2012;258(24):975661.
[16] Ikawa N, Shimada S, Tanaka H, Ohmori G. An atomistic analysis of nanometric
chip removal as affected by toolwork interaction in diamond turning. CIRP
Ann: Manuf Technol 1991;40(1):5514.
[17] Pramanik A, Neo K, Rahman M, Li X, Sawa M, Maeda Y. Ultra-precision turning of
electroless-nickel: effect of phosphorus contents, depth-of-cut and rake angle.
J Mater Proc Technol 2008;208(1):4008.
[18] Yan J, Zhang Z, Kuriyagawa T. Mechanism for material removal in diamond turning of reaction-bonded silicon carbide. Int J Mach Tools Manuf
2009;49(5):36674.
[19] Ghisleni R, Rzepiejewska-Malyska K, Philippe L, Schwaller P, Machler J. In situ
SEM indentation experiments: instruments, methodology, and applications.
Microsc Res Tech 2009;72(3):2429.
[20] Moser B, Lfer J, Michler J. Discrete deformation in amorphous metals: an
in situ SEM indentation study. Philos Mag 2006;86(3335):571528.
[21] Yu Q, Shan Z, Li J, Huang X, Xiao L, Sun J, et al. Strong crystal size effect on
deformation twinning. Nature 2010;463(7279):3358.
[22] Gao W, Hocken RJ, Patten JA, Lovingood J, Lucca DA. Construction and testing
of a nanomachining instrument. Precis Eng 2000;24(4):3208.
[23] Fatikow S, Wich T, Hulsen H, Sievers T, Jahnisch M. Microrobot system for automatic nanohandling inside a scanning electron microscope. IEEE/ASME Trans
Mechatron 2007;12(3):24452.
[24] Jasper D, Diederichs C, Edeler C, Fatikow S. High-speed nanorobot position control inside a scanning electron microscope. In: International Conference on
Electrical Engineering/Electronics Computer Telecommunications and Information Technology (ECTI-CON). 2010. p. 5137.
[25] Xu Z, Fang F, Zhang S, Hu X, Fu Y, Li L. Fabrication of micro DOE using micro
tools shaped with focused ion beam. Opt Express 2010;18(8):802532.
[26] Kawasegi N, Niwata T, Morita N, Nishimura K, Sasaoka H. Improving machining
performance of single-crystal diamond tools irradiated by a focused ion beam.
Precis Eng 2014;38(1):17482.

You might also like