Вы находитесь на странице: 1из 22

Bhushan, B. Microscratching/Microwear, Nanofabrication/Nanomachining... Handbook of Micro/Nanotribology. Ed.

Bharat Bhushan Boca Raton: CRC Press LLC, 1999

1999 by CRC Press LLC

Microscratching/ Microwear, Nanofabrication/ Nanomachining, and Nano/Picoindentation Using Atomic Force Microscopy
Bharat Bhushan
7.1 7.2 7.3 7.4 7.5 7.6 Introduction Experimental Techniques
AFM for Microscratching/Microwear and Nanoindentation Nano/Picoindenter

Microscratching/Microwear Studies Nanofabrication/Nanomachining Studies Nano/Picoindentation Closure References

7.1 Introduction
Wear of sliding surfaces can occur by one or more wear mechanisms, including adhesive, abrasive, fatigue, impact, corrosive, and fretting. The wear rate, a measure of wear, generally needs to be minimized (Bhushan, 1996). As the dimensions of components and loads used continue to decrease (such as in microelectromechanical systems or MEMS), scratching/wear and mechanical properties at the micro- to nanoscales become very important. With the advent of the newly developed scanning probe microscopes (SPMs), particularly the atomic force microscope (AFM), it is possible to study the interfacial phenomena at a small scale and light load. The AFM/FFM tip simulates a sharp single asperity traveling over a surface. Scratching and wear processes at different normal loads are studied in AFMs by using a sharp diamond tip. This tip can also be used for nanofabrication/nanomachining. AFMs, in conjunction with special sensors, are used for measurement of mechanical properties on nano- to picoscales.

1999 by CRC Press LLC

This chapter presents an overview of microscratching/microwear, nanofabrication/nanomachining, and nano/picoindentation using AFM and related instrumentation.

7.2 Experimental Techniques


7.2.1 AFM for Microscratching/Microwear and Nanoindentation Commercial AFMs are commonly used to conduct microscratching/microwear and nanoindentation. Special sensors may be used in conjunction with AFMs for nano/picoindentation studies. For microscale scratching, microscale wear, nanoscale indentation hardness measurements, and nanofabrication/nanomachining, a three-sided pyramidal single-crystal natural diamond tip with an apex angle of 80 and a radius of about 100 nm mounted on a stainless steel cantilever beam with high normal stiffness of about 25 N/m is used at relatively high loads (1 to 150 N); see Chapter 1 for further details (Bhushan et al., 1994a). For scratching and wear studies, the sample is generally scanned in a direction orthogonal to the long axis of the cantilever beam (typically at a rate of 0.5 Hz) so that friction force can be measured during scratching and wear. The tip is mounted on the beam such that one of its edges is orthogonal to the long axis of the beam; therefore, wear during scanning along the beam axis is higher (about 2 to 3 times) than that during scanning orthogonal to the beam axis. For wear studies, typically an area of 2 2 m is scanned at various normal loads (ranging from 1 to 100 N) for a selected number of cycles. For nanofabrication/nanomachining, the nanoscratching operation is extended. For nanoindentation hardness measurements, the scan size is set to zero and normal load is applied to make the indents (Bhushan et al., 1994a,d). During this procedure, the diamond tip is continuously pressed against the sample surface for about 2 s at various indentation loads. Nanohardness is calculated by dividing the indentation load by the projected residual area of the indents. Sample surface is scanned before and after the scratching, wear, or indentation to obtain the initial and the nal surface topography, at a low normal load of about 0.3 N using the same diamond tip. An area larger than the scratching, wear, and indentation region is scanned to observe the marks.

7.2.2 Nano/Picoindenter
As stated earlier, conventional AFMs have been used for indentation studies on nanometer-scale depths. In these studies, the hardness value is based on the projected residual area after imaging the indent. Direct imaging of the indent allows one to quantify piling up of ductile material around the indenter. However, it becomes difcult to identify the boundary of the indentation mark with great accuracy. This makes the direct measurement of the contact area somewhat inaccurate (Bhushan et al., 1994a,d). A technique with the dual capability of depth sensing as well as in situ imaging is most appropriate in nanomechanical property studies (Bhushan et al., 1996). This indentation system is used to make loaddisplacement measurement and subsequently carry out in situ imaging of the indent. A schematic of the nano/picoindenter system used is shown in Figure 7.1. The indentation system consists of a three-plate transducer with electrostatic actuation hardware used for direct application of normal load and a capacitive sensor used for measurement of vertical displacement. The AFM head is replaced with this transducer assembly while the specimen is mounted on the piezoelectric scanner which remains stationary during indentation experiments. The transducer consists of a three (BeCu) plate capacitive structure which provides high sensitivity, large dynamic range, and a linear output signal with respect to load or displacement. The tip is mounted on the center plate. The upper and lower plates serve as drive electrodes. Load is applied by applying appropriate voltage to the drive electrodes, thereby generating an electrostatic force between the center plate and the drive electrodes. Vertical displacement of the tip (indentation depth) is measured by measuring the displacement of the center plate relative to the two outer electrodes using capacitance technique. The load resolution is 100 nN or better, and the displacement resolution is 0.1 nm. At present, a load range of 1 N to 10 mN can be employed. Loading rates can be varied by changing the load/unload period. The AFM functions as the platform providing an in situ image of the indent with a lateral resolution of 1 nm and a vertical resolution of 0.2 nm. The loaddisplacement data

1999 by CRC Press LLC

can be acquired and displayed on the display monitor. Hardness value can be obtained from the loaddisplacement data, as well as from direct measurement of the projected residual area of the indent after imaging. Youngs modulus of elasticity is obtained from the slope of the unloading curve. A three-sided Berkovich indenter with tip radius of about 100 nm is generally used for the measurements, see Chapter 10 on Nanomechanical Properties in this book. Sharper diamond tips with included angle of 60 to 90 and tip radii of 30 to 60 nm are sometimes employed for shallower indentation (on the order of 1 nm). To obtain an accurate relation between the indentation depth and the projected contact area, tip shape calibration needs to be done. Also for surfaces with rms roughness on the order of indentation depth, the original (unindented) prole is subtracted from the indented prole (Bhushan et al., 1994a,d). In a typical indentation experiment, the tip is lowered close to the sample (ideally <100 m). Scan size and scan rate are selected. The tip is engaged to the sample surface by a stepper motor with a set point of 1 nA (about 1 N). A desired image area is captured prior to indentation. The feedback is set to zero to disable the scanner; the scan size is set to zero so that the indenter will be positioned at the center of the image. An appropriate set point for the preload condition is selected. The indentation rate can be varied by changing the load/unload period.

7.3 Microscratching/Microwear Studies


By using a standard or sharp diamond tip mounted on a stiff cantilever beam, AFMs can be used to investigate how surface materials can be moved or removed on micro- to nanoscales, for example, in scratching and wear (where these things are undesirable) and nanofabrication/nanomachining (where they are desirable) (Hamada and Kaneko, 1992; Miyamoto et al., 1991; Bhushan and Koinkar, 1994b, 1995ac,e, 1997; Bhushan et al., 1994 a,c; 1995d; Bhushan, 1995a,b, 1997, 1998a,b; Koinkar and Bhushan, 1997a,b). A variety of polymers and ceramics and hard coatings have been studied. Many examples of scratching/wear of magnetic recording materials have been presented in Chapter 14 on magnetic storage devices in this book. This chapter focuses on the studies with silicon material. Figure 7.2a shows microscratches made on Si(111) at various loads (Bhushan and Koinkar, 1994b). As expected, the scratch depth increases linearly with load. Such microscratching measurements can be used to study failure mechanisms on the microscale and to evaluate the mechanical integrity (scratch resistance) of ultrathin lms at low loads. To study the effect of scanning velocity, unidirectional scratches, 5 m in length, were generated at scanning velocities ranging from 1 to 100 m/s at normal loads ranging from 40 to 140 N. There is no effect of scanning velocity obtained at a given normal load. For representative scratches proles at 80 N, see Figure 7.2b (Koinkar, 1997). Insensitivity to scanning velocity may be because of a small effect of frictional heating with the change in scanning velocity used here. Furthermore, for a small change in interface temperature, there is a large underlying volume to dissipate the heat generated during scratching (Bhushan, 1998a). By scanning the sample in two dimensions with the AFM, wear scars are generated on the surface. Figure 7.3 shows the effect of normal load on the wear rate. We note that wear rate is very small below 20 N of normal load. A normal load of 20 N corresponds to contact stresses comparable to the hardness of the silicon. Primarily, elastic deformation at loads below 20 N is responsible for low wear (Bhushan et al., 1995d; Bhushan and Kulkarni, 1995f; Koinkar and Bhushan, 1997b). Typical wear mark generated at a normal load of 40 N for one scan cycle and imaged using AFM at 300 nN load is shown in Figure 7.4a (Koinkar and Bhushan, 1997b). The inverted map of a wear mark shown in Figure 7.4b indicates the uniform material removal at the bottom of the wear mark. Next we examine the mechanism of material removal on microscale at low loads, in AFM wear experiments. Figure 7.5 shows a secondary electron image of a wear mark and associated wear particles. The specimen used for the SEM was not scanned after initial wear, to retain wear debris in the wear region. Wear debris is clearly observed. An AFM image of the wear mark shows small debris at the edges, swiped during AFM scanning. Thus, the debris is loose (not sticky) and can be removed during the AFM scanning. SEM micrographs show both cutting-type and ribbonlike debris. TEM studies were performed to understand

1999 by CRC Press LLC

1999 by CRC Press LLC

FIGURE 7.1 Schematics of (a) indentation system, (b) three-plate transducer with electrostatic actuation hardware and capacitance sensor, and (c) tip-holder mount assembly. (From Bhushan, B. et al. (1996). Philos. Mag., 74, 1171128. With permission.)

1999 by CRC Press LLC

FIGURE 7.1

FIGURE 7.2 Surface plots of (a) Si(111) scratched for ten cycles at various loads and a scanning velocity of 2 m/s (note that the x- and y-axes are in m and the z-axis is in nm), and (b) Si(100) scratched in one unidirectional scan cycle at a normal load of 80 N and different scanning velocities.

the material removal process. The TEM micrograph of the worn region in Figure 7.6 shows evidence of bend contours passing through the wear mark. The bend contours around and inside the wear mark suggest that there are some residual stresses around and inside the wear mark region. There is no dislocation activity or cracks observed inside the wear track. The dislocation activity and/or cracking probably occurs at the subsurface. Based on SEM and TEM studies, it is believed that the material in the experiment described here is removed in a brittle manner without much plastic deformation (dislocation activity). Finally, we study evolution of wear of a diamondlike carbon (DLC) coated disk substrate, Figure 7.7 (Bhushan et al., 1994a). The data illustrate how the microwear prole for a load of 20 N develops as a function of the number of scanning cycles. Wear is not uniform, but is initiated at the nanoscratches indicating that the nanoscratches (with high surface energy) and nonuniform coverage of DLC at nanoscratches act as initiation sites. Thus, scratch-free surfaces will be relatively resistant to wear.

1999 by CRC Press LLC

FIGURE 7.3 Wear depth as a function of normal load for Si(111) after one cycle. (From Koinkar, V. N. and Bhushan, B. (1997), J. Mater. Res., 12, 32193224. With permission.)

7.4 Nanofabrication/Nanomachining Studies


Scanning tunneling microscopes (STMs) have been used to form nanofeatures by localized heating or by inducing chemical reactions under the STM tip (Abraham et al., 1986; Silver et al., 1987; Albrecht et al., 1989; Utsugi, 1990; Kobayashi et al., 1993), and nanomachining (Parkinson, 1990). AFMs have also been used for nanofabrication (Majumdar et al., 1992; Bhushan et al., 1994a,c; Bhushan, 1995a,b, 1998a,b; Tsau et al., 1994) and nanomachining (Delawski and Parkinson, 1992). Figure 7.8 shows an example of nanofabrication. The word OHIO was written on a (100) singlecrystal silicon wafer by scratching the sample surface with a diamond tip at specied locations and scratching angles (Bhushan, 1995a). The normal load used for scratching (writing) was 50 N and the writing speed was 0.2 m/s. Each line is scribed manually and debris at the ends of each line is visible. A few lines are not connected to each other because of the PZT drift and hysteresis. Sufcient time should be given for the thermal stabilization of the PZT scanner so that the hysteresis effect is small during the nanofabrication. Next, more complex patterns were generated at a normal load of 15 N and a writing speed of 0.5 m/s, Figure 7.9 (Koinkar, 1997). Such a type of patterns is useful for resistor trimming (to increase the path resistor) on a small scale. The separation between lines is about 50 nm. In Figure 7.9a, the variation in line width is due to the tip asymmetry. A spiral pattern generated as shown in Figure 7.9b. Nanofabrication parameters normal load, scanning speed, and tip geometry can be controlled precisely to control depth and length of the devices. Nanofabrication using mechanical scratching has several advantages over other techniques (Koinkar, 1997). Better control over the applied normal load, scan size, and scanning speed can be used for nanofabrication of devices. Using this technique, nanofabrication can be performed on any engineering surface. Use of chemical etching or reactions is not required, and this dry nanofabrication process can be used where use of chemicals and electric eld is prohibited. One disadvantage of this technique is the formation of debris during scratching. At light loads, debris formation is not a problem compared with high-load scratching. However, debris can be removed easily out of the scan area at light loads during scanning.

7.5 Nano/Picoindentation
Nanohardness measurements using conventional AFMs is covered in Chapter 14 on magnetic storage devices. In this chapter, we will limit the discussion to the application of the three-plate transducer with

1999 by CRC Press LLC

FIGURE 7.4 (a) Typical gray scale and (b) inverted AFM images of a wear mark created using a diamond tip at a normal load of 40 N and one scan cycle on Si(111) surface. (From Koinkar, V. N. and Bhushan, B. (1997), J. Mater. Res., 12, 32193224. With permission.)

electrostatic actuation hardware used in conjunction with conventional AFMs (Bhushan et al., 1996; Kulkarni et al., 1996a,b, 1997; Bhushan, 1997, 1998a,b; Bhushan and Koinkar, 1997; and Koinkar and Bhushan, 1997a). Figure 7.10a shows the loaddisplacement curves at different peak loads for Si(100). Loaddisplacement data at residual depth as low as about 1 nm can be obtained. Loading/unloading curves are not smooth, but exhibit sharp discontinuities particularly at high loads (shown by arrows in the gure). Any discontinuities in the loading part of the curve probably result from slip of the tip. The sharp discontinuities in the unloading part of the curves are believed to be due to formation of lateral cracks that form at the base of median crack, which results in the surface of the specimen being thrust upward. From the loaddisplacement curves in Figure 7.10 the indentation hardness of surface lms with an indentation depth of as small as about 1 nm has been measured. Triangular indentations are observed for shallow penetration depths, Figure 7.10b.

1999 by CRC Press LLC

FIGURE 7.5 Secondary electron image of wear mark and debris for Si(111) produced at a normal load of 40 N and one scan cycle. (From Koinkar, V. N. and Bhushan, B. (1997), J. Mater. Res., 12, 32193224. With permission.)

FIGURE 7.6 Bright-eld TEM micrograph showing wear mark and bend contour around and inside the wear mark in Si(111) produced at a normal load of 40 N and one scan cycle. (From Koinkar, V. N. and Bhushan, B. (1997), J. Mater. Res., 12, 32193224. With permission.)

Figure 7.11 shows the loaddisplacement curves during three loading and unloading cycles for singlecrystal silicon. The unloading and reloading curves reveal a large hysteresis, which shows no sign of degeneration through three cycles of deformation and the peak load displacement shift to higher values in successive loadingunloading cycles. Pharr et al. (1989, 1990), Page et al. (1992), and Pharr (1992) have also observed hysteresis behavior in silicon at similar loads using a nanoindenter. The fact that the curves are highly hysteretic implies that deformation is not entirely elastic. Pharr (1992) concluded that large hysteresis is due to a pressure-induced phase transformation from its normal diamond cubic form to a -tin metal phase. Table 7.1 summarizes the hardness and Youngs modulus of eleasticity data at various depths for singlecrystal silicon (Bhushan et al., 1996). Comparison of nanohardness values with that of bulk hardness

1999 by CRC Press LLC

FIGURE 7.7 Surface plots of DLC-coated thin-lm disk showing the worn region; the normal load and number of test cycles are indicated. (Bhushan, B. et al. (1994), Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol. 208, 1729. With permission.)

1999 by CRC Press LLC

FIGURE 7.8 Example of nanofabrication. The letters OHIO (which stands for the state of Ohio) were generated by scratching a Si(111) surface using a diamond tip at a normal load of 50 N and writing speed of 0.2 m/s. (Bhushan, 1995b).

shows that nanohardness of silicon is higher that its bulk hardness. Figure 7.12 shows the plot of hardness and Youngs modulus of elasticity as a function of residual depth. Note that hardness increases with a decrease in the residual depth. Similar results on single-crystal silicon have been reported by Pethica et al. (1983) and Page et al. (1992) based on nanoindentation data obtained using commercial nanoindentors, and by Bhushan and Koinkar (1994d) based on AFM data. Gane and Cox (1970) reported similar results on gold based on microindenter data. This increase in hardness with a decrease in indentation depth can be rationalized on the basis that, as the volume of deformed material increases, there is a higher probability of encountering material defects (Bhushan and Koinkar, 1995d). The increase in hardness also could be due to surface-localized cold work resulting form polishing. The nano/picoindentation system has been used to study the creep and strain rate effects of ceramics. Most materials, including ceramics and even diamond, are known to creep at temperatures well below half their melting points, even at room temperature. Indentation creep and strain rate sensitivity experiments were conducted on single-crystal silicon. Figure 7.13 shows the loaddisplacement curves for various peak loads held at 180 s. To demonstrate the creep effects, the loaddisplacement curves for a 500-N peak load held at 0 and 30 s are also shown as an inset. Note that signicant creep occurs at room temperature. Nanoindenter experiments conducted by Li et al. (1991) exhibited signicant creep only at high temperatures (greater than or equal to 0.25 times the melting point of silicon). The mechanism of dislocation glide plasticity is believed to dominate the indentation creep process. To study strain rate sensitivity of silicon, experiments were conducted at two different (constant) rates of loading (Figure 7.14). Note that a change in the loading rate of a factor of about ve results in a change in the loaddisplacement data. Strain rate sensitivity to single-crystal aluminum (111) has been reported by LaFontaine et al. (1990). Hardness of 5-nm- to 100-nm-thick DLC (ac:H) lms have been measured by Kulkarni and Bhushan (1997) and Koinkar and Bhushan (1997a). Figure 7.15a shows the loaddisplacement plots of silicon (100) and various 20-nm coatings at 100-N peak load. Peak indentation depth for silicon (100) at 100-N load is 8.7 nm and those for sputtered, ion beam and cathodic arc carbon coatings are 7.8, 8.2, and 6.4 nm, respectively. The residual depth is less than 1 nm for cathodic arc carbon coating at this load. From the comparison of these coatings just on the basis of loaddisplacement data, cathodic arc carbon exhibits the least displacement at the peak load. Representative loaddisplacement plots for

1999 by CRC Press LLC

FIGURE 7.9 (a) Trim and (b) spiral patterns generated by scratching a Si(100) surface using a diamond tip at a normal load of 15 N and writing speed of 0.5 m/s.

100-nm-thick carbon coatings at peak indentation load of 500 N are shown in Figure 7.15b. Note again that the cathodic arc carbon specimen exhibits the lowest indentation depth, about 20 nm, at the peak load as compared with other coatings. Figure 7.16a shows the plot of hardness as a function of residual depth for silicon(100) and various carbon coating. The indentation loads used in each case ranged from 100 to 2000 N. The cathodic arc carbon coating exhibits the highest hardness of about 24.9 GPa, whereas the sputtered and ion beam carbon coatings exhibit hardness values of 17.2 and 15.2 GPa, respectively. The hardness of silicon(100) is 13.2 GPa. The elastic modulus as a function of residual depth is plotted in Figure 7.16b. The cathodic arc carbon coating exhibits a decrease in the elastic modulus with increasing residual depth, while the elastic moduli for the other carbon coatings remain almost constant. In general, hardness and the elastic modulus of coatings are strongly inuenced by their crystalline structure, stoichiometry, and growth characteristics, which usually depend on the deposition parameters. Mechanical properties of aC:H coatings have been known to change over a wide range with sp3 to sp2 bonding ratio and amount of hydrogen. Hydrogen is believed to play an important role in the bonding conguration of carbon atoms by helping to stabilize tetrahedral coordination of carbon atoms.

1999 by CRC Press LLC

7.6 Closure
AFM with suitable diamond tips has been used successfully to measure microscratching/microwear and nano/picoindentation behavior of solid surfaces and thin lms. AFM has also been used for nanofabrication/nanomachining purposes. Scratch and wear properties of single-crystal silicon have been measured. Scratch and wear depths increase with an increase of normal load. Wear rate for single-crystal silicon is found to be negligible below 20 N, is much higher at higher loads, and increases almost linearly with load. Elastic deformation at low loads is responsible for negligible wear. The mechanism of material removal on the microscale is studied. At the loads used in the study, material is removed by the plowing mode in a brittle manner without much plastic deformation. Most of the wear debris is loose. Evolution of the wear of thin DLC coatings has also been studied using AFM. Wear is found to be initiated at nanoscratches. AFM has also been demonstrated to be useful for nanofabrication. AFM has been modied to obtain loaddisplacement curves and for measurement of nanoindentation hardness and Youngs modulus of elasticity, with depth of indentation as low as 1 nm. Hardness of ceramics on nanoscales is found to be higher than that on the microscale. Ceramics exhibit signicant plasticity and creep on the nanoscale. Scratching and indentation on nanoscales are the powerful ways to screen for adhesion and resistance to deformation of ultrathin lms. These studies are also directly applicable to interfacial phenomena of MEMS devices.

1999 by CRC Press LLC

(a)

FIGURE 7.10

1999 by CRC Press LLC

(b)

FIGURE 7.10 (a) Loaddisplacement curves at various peak loads for Si(100). Arrows indicate the discontinuities in the displacement; (b) AFM gray scale images of the indentations made at these peak loads. (Kulkarni, A. V. and Bhushan, B. (1997), J. Mater. Res., 12, 27072714. With permission.)

1999 by CRC Press LLC

FIGURE 7.11 Loaddisplacement curves during repeated loadingunloading cycles for Si(100) (Bhushan, B. et al. (1996), Philos. Mag., 74, 11171128. With permission.) TABLE 7.1 Comparison of Nanohardness and Youngs Modulus of Elasticity of Single-Crystal Silicon (100) as a Function of Normal Load
Normal Load (N) 500 700 1000 1200 2000
a

Residual Depth (nm) 2.8 7.0 8.5 9.5 15.0

Contact Depth (nm) 17.4 25.4 32.2 37.0 55.3

Hardnessa (GPa) 13.0 12.7 13.2 13.4 13.1

Youngs Modulus of Elasticitya 160 141 143 141 141

Bulk values of hardness and Youngs modulus of elasticity are 9 to 10 and 130 GPa, respectively.

FIGURE 7.12 Indentation hardness as a function of residual indentation depth for Si(100). (Bhushan, B. et al. (1996), Philos. Mag., 74, 11171128. With permission.)

1999 by CRC Press LLC

FIGURE 7.13 permission.)

Creep behavior of single-crystal silicon. (Bhushan, B. et al. (1996), Philos. Mag., 74, 11171128. With

FIGURE 7.14 Strain rate sensitivity of single-crystal silicon. (Bhushan, B. et al. (1996), Philos. Mag., 74, 11171128. With permission.)

1999 by CRC Press LLC

(a)

(b)

FIGURE 7.15 Loaddisplacement curves for (a) Si(100) and various 20-nm-thick DLC coatings at peak indentation load of 100 N and (b) for Si(100) and 100-nm-thick DLC coatings at peak indentation load of 500 N. Arrow indicates the discontinuity in the displacement. (Kulkarni, A. V. and Bhushan, B. (1997), J. Mater. Res. 12, 27072714. With permission.)

1999 by CRC Press LLC

FIGURE 7.16 (a) Hardness and (b) Youngs modulus of elasticity as a function of residual depth for silicon (100) and various 100-nm-thick DLC coatings. (Kulkarni, A. V. and Bhushan, B. (1997), J. Mater. Res. 12, 27072714. With permission.)

References
Abraham, D.W., Mamin, H.J., Ganz, E., and Clark, J. (1986), Surface Modication with the Scanning Tunneling Microscope, IBM J. Res. Dev. 30, 492499. Albrecht, T.R., Dovek, M.M., Kirk, M.D., Lang, C.A., Quate, C.F., and Smith, D.P.E. (1989), NanometerScale Hole Formation on Graphite Using a Scanning Tunneling Microscope, Appl. Phys. Lett. 55, 17271729. Bhushan, B. (1995a), Nanotribology and Its Applications to Magnetic Storage Devices and MEMS, Forces in Scanning Probe Methods (H.J. Guntherodt, D. Anselmetti, and E. Meyer, eds.), Vol. E 286, pp. 367395, Kluwer Academic, Dordrecht, Netherlands. Bhushan, B. (1995b), Micro/Nanotribology and Its Applications to Magnetic Storage Devices and MEMS, Tribol. Int. 28, 8595. Bhushan, B. (1996), Tribology and Mechanics of Magnetic Storage Devices, 2nd ed., Springer-Verlag, New York. Bhushan, B. (1997), Micro/Nanotribology and Its Application, Vol. E330, Kluwer Academic, Dordrecht, Netherlands.

1999 by CRC Press LLC

Bhushan, B. (1998a), Micro/Nanotribology Using Atomic Force/ Friction Force Microscopy: State of the Art, Proc. Inst. Mech Eng., Part J: J. Eng. Tribol. 212, 1-18. Bhushan, B. (1998b), Tribology Issues and Opportunities in MEMS, Kluwer Academic, Dordrecht, The Netherlands. Bhushan, B., Koinkar, V.N., and Ruan, J. (1994a), Microtribology of Magnetic Media, Proc. Inst. Mech Eng., Part J: J. Eng. Tribol. 208, 1729. Bhushan, B. and Koinkar, V.N. (1994b), Tribological Studies of Silicon for Magnetic Recording Applications, J. Appl. Phys. 75, 57415746. Bhushan, B., Koinkar, V.N., and Ruan, J. (1994c), Microtribological Studies by Using Atomic Force and Friction Force Microscopy and Its Applications, in Determining Nanoscale Physical Properties of Materials by Microscopy and Spectroscopy (M. Sarikaya, H.K. Wickramasinghe and M. Isaacson, eds.), Vol. 332, pp. 9398, Materials Research Society, Pittsburgh. Bhushan, B. and Koinkar, V.N. (1994d), Nanoindentation Hardness Measurements Using Atomic Force Microscopy, Appl. Phys. Lett. 64, 16531655. Bhushan, B. and Koinkar, V.N. (1995a), Microtribology of PET Polymeric Films, Trib. Trans. 38, 119127. Bhushan, B. and Koinkar, V.N. (1995b), Macro and Microtribological Studies of CrO2 Video Tapes, Wear 180, 916. Bhushan, B. and Koinkar, V.N. (1995c), Microtribology of Metal Particle, Barium Ferrite and Metal Evaporated Magnetic Tapes, Wear 183, 360370. Bhushan, B., Israelachvili, J.N., and Landman, U. (1995d), Nanotribology: Friction, Wear and Lubrication at the Atomic Scale, Nature 374, 607616. Bhushan, B. and Koinkar, V.N. (1995e), Microscale Mechanical and Tribological Characterization of Hard Amorphous Carbon Coatings as Thin as 5 nm for Magnetic Disks, Surf. Coatings Technol. 7677, 655669. Bhushan, B., and Kulkarni, A.V. (1995f), Effect of Normal Load on Microscale Friction Measurements, Thin Solid Films 278, 4956. Bhushan, B., Kulkarni, A.V., Bonin, W., and Wyrobek, J.T. (1996), Nano/Picoindentation Measurement Using a Capacitive Transducer System in Atomic Force Microscopy, Philos. Mag. 74, 11171128. Bhushan, B., and Koinkar, V.N. (1997), Microtribological Studies of Doped Single-Crystal Silicon and Polysilicon Films for MEMS Devices, Sensors Actuators A 57, 91102. Delawski, E. and Parkinson, B.A. (1992), Layer-by-Layer Etching of Two-Dimensional Metal Chalcogenides with the Atomic Force Microscope, J. Am. Chem. Soc. 114, 16611667. Gane, N. and Cox, J.M. (1970), The Micro-Hardness of Metals at Very Light Loads, Philos. Mag. 22, 881891. Hamada, E. and Kaneko, R. (1992), Microdistortion of Polymer Surfaces by Friction, J. Phys. D: Appl. Phys. 25, A53A56. Kobayashi, A., Grey, F., Williams, R.S., and Ano, M. (1993), Formation of Nanometer-Scale Grooves in Silicon with a Scanning Tunneling Microscope, Science 249, 17241726. Koinkar, V.N. (1997), Micro/Nanotribology and Its Applications to Magnetic Media, Heads and MEMS, Ph.D. Dissertation, The Ohio State University, Columbus. Koinkar, V.N., and Bhushan, B. (1997a), Microtribological Properties of Hard Amorphous Carbon Protective Coatings for Thin-Film Magnetic Disks and Heads, Proc. Inst. Mech. Eng. Part J: J. Eng. Tribol. 211, 365372. Koinkar, V.N. and Bhushan, B. (1997b), Scanning and Transmission Electron Microscopies of SingleCrystal Silicon Microworn/Machined Using Atomic Force Microscopy, J. Mater. Res. 12, 32193224. Kulkarni, A.V. and Bhushan, B. (1996a), Nanoscale Mechanical Property Measurements Using Modied Atomic Force Microscopy, Thin Solid Films 290291, 206210.

1999 by CRC Press LLC

Kulkarni, A.V. and Bhushan, B. (1996b), Nano/Picoindentation Measurements on Single-Crystal Aluminum Using Modied Atomic Force Microscopy, Mater. Lett. 29, 221227. Kulkarni, A.V. and Bhushan, B. (1997), Nanoindentation Measurements of Amorphous-Carbon Coatings, J. Mater. Res., 12, 27072714. LaFontaine, W.R., Yost, B., Black, R.D., and Li, C.Y. (1990), Indentation Load Relaxation Experiments with Indentation Depth in the Submicron Range, J. Mater. Res. 5, 21002106. Li, W.B., Henshall, J.L., Hooper, R.M., and Easterling, K.E. (1991), The Mechanism of Indentation Creep, Acta Metall. Mater. 39, 30993110. Majumder, A., Oden, P.I., Carrejo, J.P., Nagahara, L.A., Graham, J.J., and Alexander, J. (1992), Nanometer-Scale Lithography Using the Atomic Force Microscope, Appl. Phys. Lett. 61, 22932295. Miyamoto, T., Kaneko, R., and Miyake, S. (1991), Tribological Characteristics of Amorphous Carbon Films Investigated by Point Contact Microscopy, J. Vac. Sci. Technol. B 9, 13381339. Page, T.F., Oliver, W.C., and McHargue, C.J. (1992), The Deformation Behavior of Ceramic Crystals Subjected to Very Low Load (Nano) Indentations, J. Mater. Res. 7, 450473. Parkinson, B. (1990), Layer-by-Layer Nanometer Scale Etching of Two-Dimensional Substrates Using the Scanning Tunneling Microscopy, J. Am. Chem. Soc. 112, 74987502. Pethica, J.N., Hutchings, R., and Oliver, W.C. (1983), Hardness Measurements at Penetration Depths as Small as 20 nm, Philos. Mag. A 48, 598606. Pharr, G.M. (1992), The Anomalous Behavior of Silicon During Nanoindentation, in Thin Films: Stresses and Mechanical Properties, III (W.D. Nix, J.C. Braveman, E. Arzt, and L.B. Freund, eds.), Vol. 239, pp. 301312, MRS, Pittsburgh. Pharr, G.M., Oliver, W.C., and Clarke, D.R. (1989), Hysteresis and Discontinuity in the Indentation Load-Displacement Behavior of Silicon, Scrip. Metall. 23, 19491952. Pharr, G.M., Oliver, W.C., and Clarke, D.R. (1990), The Mechanical Behavior of Silicon during SmallScale Indentation, J. Electr. Mater. 19, 881887. Silver, R.M., Ehrichs, E.E., and deLozanne, A.L. (1987), Direct Writing of Submicron Metallic Features with a Resonance, Phys. Rev. Lett. 70, 35063509. Tsau, L., Wang, D., and Wang, K. L. (1994), Nanometer Scale Patterning of Silicon(100) Surface by an Atomic Force Microscope Operative in Air, Appl. Phys. Lett. 64, 21332135. Utsugi, Y. (1990), Nanometer-Scale Chemical Modication Using a Scanning Tunneling Microscope, Nature 347, 747749.

1999 by CRC Press LLC

Вам также может понравиться