You are on page 1of 9

A Fresh Insight Into the

Microcantilever-Sample
Interaction Problem in
Nader Jalili* Non-Contact Atomic Force
Mohsen Dadfarnia Microscopy
Smart Structures and NEMS Laboratory, The atomic force microscope (AFM) system has evolved into a useful tool for direct
Department of Mechanical Engineering, measurements of intermolecular forces with atomic-resolution characterization that can
Clemson University, Clemson, SC 29634 be employed in a broad spectrum of applications. The non-contact AFM offers unique
advantages over other contemporary scanning probe techniques such as contact AFM and
scanning tunneling microscopy, especially when utilized for reliable measurements of soft
Darren M. Dawson samples (e.g., biological species). Current AFM imaging techniques are often based on a
Robotics and Mechatronics Laboratory, lumped-parameters model and ordinary differential equation (ODE) representation of the
Department of Electrical and Computer micro-cantilevers coupled with an adhoc method for atomic interaction force estimation
Engineering, (especially in non-contact mode). Since the magnitude of the interaction force lies within
Clemson University, Clemson, SC 29634 the range of nano-Newtons to pica-Newtons, precise estimation of the atomic force is
crucial for accurate topographical imaging. In contrast to the previously utilized lumped
modeling methods, this paper aims at improving current AFM measurement technique
through developing a general distributed-parameters base modeling approach that re-
veals greater insight into the fundamental characteristics of the microcantilever-sample
interaction. For this, the governing equations of motion are derived in the global coor-
dinates via the Hamilton’s Extended Principle. An interaction force identification scheme
is then designed based on the original infinite dimensional distributed-parameters system
which, in turn, reveals the unmeasurable distance between AFM tip and sample surface.
Numerical simulations are provided to support these claims. 关DOI: 10.1115/1.1767852兴

1 Introduction provide cantilever deflection feedback 共refer to Fig. 1兲. The prin-
ciple of AFM operation is to scan the tip over the sample surface
The atomic force microscope 共AFM兲 system has evolved into a
with feedback mechanisms that enable the piezoelectric scanners
useful tool for direct measurements of intermolecular forces with
to maintain the tip at a constant force 共to obtain height informa-
atomic-resolution characterization that can be employed in a
tion兲, or constant height above the sample surface 共to obtain force
broad spectrum of applications such as electronics, semi-
conductors, materials, manufacturing, polymers, biological analy- information兲. As the tip scans the surface of the sample, moving
sis, and biomaterials. Specifically, AFM-based systems provide up and down with the contour of the surface, the laser beam
additional capabilities and advantages relative to other micro- deflected from the cantilever provides measurements of the differ-
scopic methods 共e.g., scanning electron microscopy 共SEM兲 and ence in light intensities between the upper and lower photo detec-
transmission electron microscopy 共TEM兲兲 with regard to studies tors, and then converted to voltage. Feedback from the photodiode
of metallic surfaces and microstructures by providing reliable difference signal, through software control from the computer,
measurements at the nanometer scale 共关1– 4兴兲. AFM can also be enables the tip to maintain either a constant force or constant
used for nano-indentation to provide in situ imaging ability with- height above the sample. In the constant force mode, the piezo-
out moving the sample, switching tips, relocating the area for electric transducer monitors real time height deviation, while in
scanning, or using an entirely different instrument to image the the constant height mode, the deflection force on the sample is
indentation 共关5–7兴兲. Assembly of nanoparticles and linking them recorded.
to electrical leads, such as random deposition of clusters between 1.2 AFM Operational and Control Modes. The AFM im-
electrodes 关8兴, binding by wet chemistry 关9兴, and electrostatic aging systems typically operate in three open-loop modes: i兲 non-
trapping 关10兴, all serve as other important applications of the AFM contact mode, ii兲 contact mode, and iii兲 tapping mode. In order to
technique. probe electric, magnetic, and/or atomic forces of a selected
1.1 AFM Background and Overview. A typical AFM sys- sample, the non-contact mode is utilized by moving the cantilever
tem consists of a micro-cantilever probe with a sharp tip mounted slightly away from the sample surface and oscillating the cantile-
to a piezoelectric actuator with a position sensitive photo detector ver at or near its natural resonance frequency 共see Fig. 2兲. By
receiving a laser beam reflected off the end-point of the beam to mounting the cantilever on a piezoelectric element and measuring
the shift from its natural resonance frequency due to sample at-
*Corresponding author: phone: 共864兲 656-5642; fax: -4435; e-mail: tractive interactions, topographical information of the sample can
jalili@clemson.edu. be extracted 关11兴. Alternatively, the contact mode acquires sample
Contributed by the Dynamic Systems, Measurement, and Control Division of THE attributes by monitoring interaction forces while the cantilever tip
AMERICAN SOCIETY OF MECHANICAL ENGINEERS for publication in the ASME
JOURNAL OF DYNAMIC SYSTEMS, MEASUREMENT, AND CONTROL. Manuscript
remains in contact with the target sample 关12兴. The tapping mode
received by the ASME Dynamic Systems and Control Division May 13, 2003; final of operation combines qualities of both the contact and non-
revision, October 23, 2003. Associate Editor: R. Gao. contact modes by gleaning sample data through oscillating the

Journal of Dynamic Systems, Measurement, and Control JUNE 2004, Vol. 126 Õ 327
Copyright © 2004 by ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


forces used by contact mode, the tip must be given a small oscil-
lation so that these small forces can be detected by measuring the
change in amplitude, phase, or frequency of the oscillating canti-
lever. Current practice calls for detection of an amplitude shift that
occurs when the tip is in the Van der Waals regime. In many cases,
the fluid contaminant layer is substantially thicker than the range
of the Van der Waals force gradient and therefore, attempts to
image the true surface with NC-AFM fail as the oscillating probe
becomes trapped in the fluid layer or hovers beyond the effective
range of the forces it attempts to measure. This is a major draw-
back of the NC-AFM imaging techniques, which significantly de-
grades the resolution of the generated topographical images. A
barely contact mode in practice is used to determine a set point
共where tip touches sample兲, around which a feedback loop is uti-
lized to control the sample positioning with respect to the tip.
Another major shortfall of the NC-AFM systems is the lack of
Fig. 1 „Left… Schematic depicting basic AFM operation and measurement of the distance between sample surface and micro-
sub-components, and „right… real scale drawing cantilever tip, i.e., Z 0 in Fig. 2 共left兲. For highest resolution, it is
necessary to measure this distance which will ultimately result in
improved interaction force estimation and enhanced topographical
cantilever tip at or near its natural resonance frequency while images.
allowing the cantilever tip to impact the target sample for a mini- 1.4 AFM Modeling Approaches. Extensive research has
mal amount of time. Though widely practiced, open-loop opera- been done on modeling and characterization of AFM tip-sample
tion modes of AFM exhibit the potential for chaotic behavior in surface interaction. Current modeling practices call for simple
the cantilever tip displacement thus rendering low resolution to- lumped model where a set of ordinary differential equations
pographical information. As a result, recent research on AFM sys- 共ODEs兲 governs the system dynamics. For example, Hsu et al.
tems has focused on detailed numerical analysis such that this 关14兴 utilized an ODE based interaction model and feedback lin-
chaotic behavior region can be well defined and ideally avoided earization and singular perturbation techniques to design an out-
关11兴. In addition to analytical methodology, several feedback con- put, high-gain feedback sample surface tracking controller. Ash-
trol strategies have been developed in order to improve the AFM hab et al. 关13兴 utilized the Melnikov method to analyze the system
region operation 关13兴. dynamics and subsequently developed a proportional/derivative
1.3 Non-Contact AFM „NC-AFM… Systems. The non- based feedback strategy to inject artificial damping such that the
contact AFM 共NC-AFM兲 offers unique advantages over other con- possibility of chaotic operation is removed.
temporary scanning probe techniques such as contact AFM 共C- The distributed nature of the structural flexibility in the micro-
AFM兲 and scanning tunneling microscopy 共STM兲. Unlike STM cantilever is an issue of increasing concern due to its direct influ-
and C-AFM, the absence of repulsive forces in NC-AFM permits ence on image resolution 共i.e., reducing the accuracy of the gen-
the imaging of ‘‘soft’’ samples, and hence, provides topography erated topographical images兲. For instance, the aforementioned
with little or no contact between the tip and the sample. Since the low-order ODE approximations mask the non-minimum phase na-
samples are not contaminated or damaged through contact with ture of the cantilever used in AFM measurements. In addition,
the AFM tip, this mode of operation is especially preferred in such low-order ODE approximations are subject to spill-over
many manufacturing processes such as non-destructive surface problems which can result in instability issues. Hence, a need
texture characterization, metrology for MEMS, 3-D topography exists for a distributed-based modeling of microcantilever-sample
for nanoparticle manufacturing and molecular manufacturing. interaction problem, which is crucial for accurate topographical
In this mode, the cantilever tip hovers 50–150 Angstrom above imaging as discussed next.
the sample surface to detect the attractive Van der Waals forces
acting between the tip and the sample, and topographic images are 2 Modeling of Atomic Force Microscopy „AFM…
constructed by scanning the tip above the surface. Since the at- There are two different types of modeling for approaches AFM;
tractive forces from the sample are substantially weaker than the lumped-parameters modeling and distributed-parameters model-

Fig. 2 Schematic diagram of non-contact AFM operation

328 Õ Vol. 126, JUNE 2004 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 4 Schematic of distributed-parameters model of the AFM
system

ics is presented here. The cantilever-sample interaction system


model of 关13兴 is extended to include the distributed nature of the
micro-cantilever assembly.
For the purpose of model development, we consider the Euler-
Bernoulli model for the microcantilever beam. As shown in Fig. 4,
Fig. 3 Lumped-parameters model of the AFM system one end of the beam is clamped into the base position assembly
with the total mass of m, while a tip mass 共here a point-mass兲,
m e , is attached to the free end of the beam. The micro-cantilever
beam has rigidity EI, linear density ␳ and the length L. We also
ing. In this section, a brief overview of the lumped-parameters consider the viscous air damping and structural damping in the
modeling approach is first presented followed by a more detailed microcantilever structure 关15兴. These types of damping terms are
modeling effort on the disturbed-parameters case. incorporated into the Euler-Bernoulli beam equation in order to
make the system exponentially stable 关16兴.
2.1 Lumped-Parameters Modeling. The lumped-para- An energy-based method is now used to derive the equations of
meters model of the atomic force microscope is shown in Fig. 3. motion. The total kinetic energy of the system can be expressed as
The dynamics for this system can be represented in the following
form 关13兴
md̈ 共 t 兲 ⫹b 共 ḋ 共 t 兲 ⫺ẋ 共 t 兲兲 ⫹k 共 d 共 t 兲 ⫺x 共 t 兲兲 ⫽ f 共 t 兲
T⫽
1
2
mḋ 2 共 t 兲 ⫹
1
2 冕 0
L
␳ 共 ḋ 共 t 兲 ⫹u t 共 x,t 兲兲 2 dx
(1)
m e ẍ 共 t 兲 ⫹b 共 ẋ 共 t 兲 ⫺ḋ 共 t 兲兲 ⫹k 共 x 共 t 兲 ⫺d 共 t 兲兲 ⫹ f IL 共 t 兲 ⫽0 1
⫹ m 共 ḋ 共 t 兲 ⫹u t 共 L,t 兲兲 2 (5)
where d(t) and x(t) denote the base motion and the cantilever tip 2 e
displacement relative to the fixed base frame coordinate, respec- where u(x,t)苸R1 represents the lateral displacement of the
tively, m苸R1 , m e 苸R1 , b苸R1 , k苸R1 denote equivalent base micro-cantilever beam respect to its base 共notice the subscripts
mass, cantilever tip mass, damping coefficient, and spring stiff- (•) t and (•) x indicate the partial derivatives with respect to the
ness coefficient, respectively, f (t)苸R1 represents the controller
time variable t and position variable x, respectively兲, d(t), ḋ(t),
force input, and f IL (t)苸R1 denotes the van der Waals attraction/
repulsion force 共i.e., the interaction forces兲 for lumped-parameters d̈(t) represent the base unit displacement 共i.e., the PZT posi-
model which is explicitly defined in the following form 关13兴 tioner兲, velocity, and acceleration signals, respectively. Neglecting
the effect of gravity, the total potential energy of the system is
Dk ␴ 6 Dk written as
f IL 共 t 兲 ⫽ ⫺

(2)
共 z 0 共 t 兲 ⫺x 共 t 兲兲 2
30共 z 0 共 t 兲 ⫺x 共 t 兲兲 8 1 L
U⫽ EIu 2xx 共 x,t 兲 dx (6)
1
where z 0 (t)苸R represents the distance from the fixed base 2 0
frame coordinate to the sample 共notice, Z 0 (t)⫽z 0 (t)⫺x(t) in Fig.
3兲, ␴ 苸R1 denotes the molecular diameter, and the model param- In the Hamiltonian approach, the boundary control input force
eter D苸R1 is defined in the following form f (t) at the base unit, the atomic interaction force f ID (t) acting at
the tip and the damping effects are all collected in the following
A HR virtual work expression
D⫽ (3)


6k L

where A H 苸R1 denotes the Hamaker constant, and R苸R1 repre- ␦ W⫽ f 共 t 兲 ␦ d 共 t 兲 ⫹ f ID 共 t 兲 ␦ u 共 L,t 兲 ⫺B u t 共 x,t 兲 ␦ u 共 x,t 兲 dx
0
sents the cantilever tip radius. Furthermore, the total cantilever tip
displacement is constrained by the following inequality
共 z 0 共 t 兲 ⫺x 共 t 兲兲 ⭓R (4)
⫺C 冕 0
L
u xt 共 x,t 兲 ␦ u 共 x,t 兲 dx (7)

2.2 Distributed-Parameters Modeling. In contrast to the where B and C are the viscous and structural damping coeffi-
previously utilized lumped modeling methods, a general cients, respectively. f ID (t), the atomic interaction force in
distributed-parameters base modeling approach that reveals distributed-parameters model, is described by the following rela-
greater insight into the fundamental characteristics of the dynam- tionship 关11兴

Journal of Dynamic Systems, Measurement, and Control JUNE 2004, Vol. 126 Õ 329

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


H1 H2 where k r and k p 苸R1 denote positive constant controller gains,
f ID 共 t 兲 ⫽ ⫺ and f f (t)苸R1 represents an auxiliary controller input. The auxil-
共 z 0 共 t 兲 ⫺u 共 L,t 兲 ⫺d 共 t 兲兲 2 30共 z 0 共 t 兲 ⫺u 共 L,t 兲 ⫺d 共 t 兲兲 8 iary control input f f (t) denotes a user-defined feedforward term
(8) utilized to operate the AFM within its three sampling regimes: 共i兲
with H 1 and H 2 representing the Hamacker constants, and z 0 (t) non-contact mode, 共ii兲 contact mode, and 共iii兲 tapping mode. In
representing the distance from the fixed base frame coordinate to the non-contact mode of operation, f f (t) in 共17兲 could be set equal
the sample 共see Fig. 4兲. The equations of motion can now be to a sinusoidal signal with a prescribed amplitude and frequency.1
obtained using the Hamilton’s principle This sinusoidal offset 共i.e., a dither signal兲 causes the AFM micro-


cantilever to oscillate about its equilibrium position. In regards to
t2
stability, f f (t) is assumed to be a bounded signal in the sense that
共 ␦ T⫺ ␦ U⫹ ␦ W 兲 dt⫽0 (9) the following inequality exists
t1

Consequently, the microcantilever governing equations can be ob- 兩 f f 共 t 兲 兩 ⭐C 4 (18)


tained as where C 4 苸R denotes a positive constant. In 关18兴, it is shown
1

that the controller given by 共17兲 ensures that all signals remain
␳ 共 u tt 共 x,t 兲 ⫹d̈ 共 t 兲兲 ⫹Bu t 共 x,t 兲 ⫹Cu xt 共 x,t 兲 ⫹EIu xxxx 共 x,t 兲 ⫽0
(10) bounded as the base of the AFM is excited by the open-loop signal

共 m⫹ ␳ L⫹m e 兲 d̈ 共 t 兲 ⫹ 冕 L

0
␳ u tt 共 x,t 兲 dx⫹m e u tt 共 L,t 兲 ⫽ f 共 t 兲 ⫹ f ID 共 t 兲
provided by the feedforward component f f (t). As further illus-
trated in 关18兴, the proposed interaction force estimator detailed
below exploits this boundedness of signal property to facilitate the
with the boundary conditions asymptotic identification of the unmeasurable time-varying inter-
action force and ultimately the distance Z 0 (t) in Fig. 2-left 共no-
u 共 0,t 兲 ⫽u x 共 0,t 兲 ⫽u xx 共 L,t 兲 ⫽0 tice, Z 0 (t)⫽z 0 (t)⫺u(L,t)⫺d(t)).
(11)
m e 关 d̈ 共 t 兲 ⫹u tt 共 L,t 兲兴 ⫺EIu xxx 共 L,t 兲 ⫽ f ID 共 t 兲 3.2 Interaction Force Estimator. In addition to ensuring
that all signals remain bounded 共i.e., the stability of the closed-
Furthermore, the total cantilever tip displacement is assumed to be loop system is preserved兲, the crucial quantity that must be esti-
constrained by the following inequality mated is the distance between sample surface and micro-
cantilever tip. That is, if one can produce a reliable estimate for
共 z 0 共 t 兲 ⫺u 共 L,t 兲 ⫺d 共 t 兲兲 ⭓Q (12) Z 0 (t), then cantilever position measurements along with Z 0 (t)
where Q denotes some positive constant; hence, the interaction can be used to generate high resolution atomic-resolution images
force f ID (t) can be upper bounded 共i.e., f ID (t)苸L ⬁ ) 共关17,18兴兲 in the non-contact mode of operation. Unfortunately, the distance
Z 0 (t) is a time-varying signal that is difficult to measure directly;
H1 H2 hence, it must be estimated by some methods. From 共8兲, it is clear
兩 f ID 兩 ⭐ ⫹ (13) that if the atomic interaction force, denoted by f ID (t), is estimated
Q2 30Q 8 and hence standard measurements 共i.e., tip and base positions兲 can
be utilized to generate an estimate of Z 0 (t). To this end, the struc-
3 Interaction Force Identification Scheme ture of the system dynamics 共10兲–共11兲 and the corresponding sta-
This section provides an identification method for the interac- bility analysis 共关17,18兴兲 motivate the following interaction force
tion force between microcantilever tip and sample surface for the estimator, denoted by f̂ ID (t)苸R1

冕 冕
distributed-parameters model of the AFM. t t
3.1 Error Dynamics System. To illustrate the technique, f̂ ID 共 t 兲 ⫽ 共 ␬ 0 ⫹1 兲 e 共 ␶ 兲 d ␶ ⫹ ␣ sgn共 e 共 ␶ 兲兲 d ␶ ⫹ 共 ␬ 0 ⫹1 兲 e 共 t 兲
we first transform the open-loop dynamics of 共10兲 and the bound- t0 t0

ary conditions 共11兲 into a more suitable structure in order to fa- ⫺ 共 ␬ 0 ⫹1 兲 e 共 t 0 兲 (19)
cilitate the appropriate design of the stabilizing boundary control
input force signal f (t). After some manipulations, it can be shown where ␬ 0 and ␣ 苸R represent positive estimator gains utilized to
1

that the open-loop boundary dynamics of 共10兲 can be rewritten in enhance the estimator performance, sgn共•兲 denotes the standard
the following more suitable form 关18兴 signum function, the free-end cantilever velocity observation error

冕 L
signal e(t)苸R1 is defined as
mṙ 0 共 t 兲 ⫽ 共 m ␤ 0 ⫺BL⫺C 兲 w t 共 0,t 兲 ⫹B w t 共 x,t 兲 dx⫹Cw t 共 L,t 兲 e 共 t 兲 ⫽w t 共 L,t 兲 ⫺ŵ t 共 L,t 兲 , (20)
0
and the dynamics for the free-end velocity observer signal ŵ t (L,t)
⫺EIw xxx 共 0,t 兲 ⫹ f 共 t 兲 (14) is given by the following
where the stabilizing error signal r 0 (t)苸R1 is defined as
m e ŵ tt 共 L,t 兲 ⫺EIw xxx 共 L,t 兲 ⫽ f̂ ID 共 t 兲 (21)
r 0 共 t 兲 ⫽w t 共 0,t 兲 ⫹ ␤ 0 w 共 0,t 兲 (15)
Based on a Lyapunov stability analysis 共see 关20兴 and 关21兴 for
with ␤ 0 苸R1 being a positive weighting constant and the auxiliary similar types of arguments兲, we can prove that the estimated in-
variable w(x,t)苸R1 defined as teraction force, denoted by f̂ ID (t), asymptotically approaches the
actual interaction force 关18兴; hence, the estimated interaction force
w 共 x,t 兲 ⫽u 共 x,t 兲 ⫹d 共 t 兲 (16) could be used to generate the required distance Z 0 (t). It should be
Since the error signal r 0 (t) is related to w(0,t) through an expo- noted that this type of estimator and its corresponding analysis is
nentially stable transfer function, we know that if r 0 (t) is bounded different from previous work in which we exploit the structure of
signal 共i.e., r 0 (t)苸L ⬁ ), then w(0,t) is also bounded signal 共i.e., the interaction force to facilitate the estimation of an unmeasur-
w(0,t)苸L ⬁ 关19兴兲. Based on the stability analysis and the structure able time-varying signal.
of 共14兲, we have developed the following control force input 关18兴 One of the primary advantages of such approach is its simplic-
applied at the base unit, ity and ease of implementation. Specifically, if we neglect shear
forces at the end of the cantilever 共these forces are being ne-
f 共 t 兲 ⫽⫺ 共 m ␤ 0 ⫺BL⫺C 兲 w t 共 0,t 兲 ⫺k r r 0 共 t 兲 ⫺k p w 共 0,t 兲 ⫹ f f 共 t 兲
(17) 1
If the frequency is set to zero, then f f (t) becomes a constant bias offset.

330 Õ Vol. 126, JUNE 2004 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


stance, from r 1 to r 2 measured from the AFM base兲, the voltage
across the thickness of the sensor can be related to microcantilever
deflection by:

v s 共 t 兲 ⫽C s 冉 ⳵x

⳵x 冊
⳵ u 共 r 1 ,t 兲 ⳵ u 共 r 2 ,t 兲
, where

C s ⫽⫺b p 共 t b /2⫹t p 兲 k 231/C p 共 r 2 ⫺r 1 兲 g 31 (22)


C p is the PZT capacitance, k 31 is PZT electromagnetic coupling
constant, t b is the microcantilever total thickness, b p is the piezo-
electric width, t p is the piezoelectric thickness, and g 31 is PZT
stress constant 共关27,28兴兲.
The electrical equivalent model of a piezoelectric element is
Fig. 5 Schematic diagram of the proposed shear force mea- composed of a capacitor with either charge generators in parallel
surement module utilizing piezoelectric film sensors or voltage sources in series. One charge 共or voltage兲 generator (q p
and v p ) accounts for the piezoelectric properties of the material,
and the capacitor represents the dielectric properties of the piezo-
glected in current AFM technology兲, then no hardware modifica- electric element. Assuming that sensor can be made very small in
tions are required with regard to a typical vendor provided AFM length 共i.e., ␧⫽r 2 ⫺r 1 is very small兲, two such sensors can be
system. The proposed force interaction estimation strategy only mounted onto the free end of the microcantilever at a very small
required free-end cantilever position measurements which are distance h from one another as shown in Fig. 5. For instance, for
readily measurable in most AFM system via the vendor mounted the left most PZT in Fig. 5, we have r 1 ⫽L⫺␧ and r 2 ⫽L. After
laser system. Hence, the proposed improved AFM imaging tech- taking into account the fact that ␧ can be made very small com-
nique could be fused with many different types of AFM systems pared to microcantilever length L, substituting these values for r 1
and AFM applications. and r 2 into 共22兲, and utilizing a backward difference derivative
approximation, we can show that voltage across the left most PZT
4 Improved AFM Imaging via Shear Measurement in Fig. 5 at about L can be approximately given by:
Module v s1 共 t 兲 ⫽⫺␧C s u xx 共 L,t 兲 . (23)
The proposed systems theory-based approach given in the pre- The voltage expression given in 共23兲 yields a real-time strain
vious sections offers fresh insight into improving NC-AFM gen- measurement at L. Similarly, the strain measurement can be ac-
erated images. Specifically, we can see from inspection of 共11兲 quired at L⫺␧⫺h for the right PZT. Consequently, the output
that the shear force 共i.e., EIw xxx (L,t)) must be distinguished from from each piezoelectric element can be sent to a numeric differ-
the atomic interaction force 共i.e., f ID (t)) if one expects to gener- entiator 共or through an advanced filtering process兲 whose output
ate accurate typographical images 共i.e., the shear forces corrupt will yield the spatial derivative of the strain in the element, and
the left-hand side of 共11兲兲. Indeed, one can see from 共21兲 that the thereby, result in the following expression for the third spatial
interaction force estimator requires shear force measurements derivative of the tip deflection used for producing shear force
共i.e., EIw xxx (L,t)) at the free-end of the microcantilever. Hence, measurements
we must develop a method for measuring the shear force at the
free-end of the microcantilever. It should be noted that shear force u xxx 共 L,t 兲 ⬵ 关 u xx 共 L 兲 ⫺u xx 共 L⫺␧⫺h 兲兴 / 共 h⫹␧ 兲 (24)
measurements have been utilized in other applications such as
After utilizing the output voltage expression for v s1 in 共23兲 and a
rotating arms 共关22–24兴兲. In these applications, two load cells are
similar voltage expression for the left PZT 共i.e., v s2 in Fig. 5兲, an
fixed at the point of interest for shear force measurement. At the
approximation of 共24兲 can be used to acquire the required shear
stationary state, there is an equal constant force acting on the two
force measurement at the tip as follows
sensors. When there is vibration, however, there will be differ-
ences between the two sensors outputs, and hence, the shear force EIw xxx 共 L,t 兲 ⫽EI 共 v s2 共 t 兲 ⫺ v s1 共 t 兲兲 /␧ 共 h⫹␧ 兲 C s (25)
at this point can be detected 关25兴.
where the relationship of 共16兲 was utilized 共i.e., w xxx (x,t)
There have been some attempts to include the shear force at
⫽u xxx (x,t)). The schematic of Fig. 5 depicts the proposed setup
microcantilever tip by adding a lateral component 共spring and
that consists of two piezoelectric elements along with the numeric
damper兲 in addition to the normal interaction force 关26兴. It appears
differentiators. The electrical equivalent model of the piezoelectric
that, however, most AFM systems neglect the corrupting influence
element and sensing mechanism are also detailed in Fig. 5 for one
of the free-end shear force. This lack of reference to this corrupt-
of the PZT element 共i.e., v s1 in Fig. 5兲. Although it is commonly
ing influence may be due to the fact that most of the literature
recognized that this type of measurement method may introduce
does not approach the AFM imaging problem by examining the
noise, our past experience with this type of noise make us believe
distributed dynamics and the associated boundary conditions. That
that this issue is not severe enough to cause insurmountable prob-
is, to make the analysis more simple, most of the previous ap-
lems 关29兴 due to the high-sampling frequency capability of mod-
proaches treat the cantilever dynamics as a lumped-parameters
ern day data acquisition hardware.
system governed by constant coefficient linear ordinary differen-
tial equations, and hence, lose much of the insight provided by the
set of dynamics given by 共10兲–共11兲. 5 Numerical Simulations and Discussion
In order to implement the estimation scheme proposed in Sec-
In order to show the improvement in the modeling, the AFM
tion 3.2, one must measure the corrupting effect of the shear force
system depicted in Fig. 4 is considered with the system param-
共see Eqs. 共19兲–共21兲兲. To measure the shear force in the cross
eters listed in Table 1. In this section, the method for solving the
section of the beam, we note that the bending strain is propor-
distributed-parameters model of AFM is explained and the results
tional to w ⬙ (x,t). Therefore, a piezoelectric film 共see Fig. 5兲 can are presented for both distributed-parameters and lumped-
be used for measuring the strain of the microcantilever, in which parameters models.
the sensor voltage due to the direct piezoelectric effect will con-
tain the strain information that can be utilized to produce the 5.1 Distributed-Parameters Model Simulations. For the
needed shear force measurements. More specifically, by bonding numerical simulations only, we utilize assumed mode model
these film sensors on the surface of the microcantilever 共for in- 共AMM兲 expansion to truncate the original partial differential

Journal of Dynamic Systems, Measurement, and Control JUNE 2004, Vol. 126 Õ 331

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 System parameters for numerical simulations ⬁

Properties Symbol Value Unit


A i d̈ 共 t 兲 ⫹N i q̈ i 共 t 兲 ⫹ 兺C
j⫽1
i j q̇ j 共 t 兲 ⫹S i q i 共 t 兲

Micro-cantilever beam rigidity EI 3.0053⫻10⫺11 Nm2


Micro-cantilever beam thickness tb 4 ␮m ⫽D 1i f ID 共 t 兲 ⫹D 2i ḟ ID 共 t 兲 ⫹D 3i f̈ ID 共 t 兲 , i⫽1,2, . . .
Micro-cantilever beam width b 35 ␮m (32b)
Micro-cantilever beam length L 125 ␮m
Micro-cantilever beam linear ␳ 3.262⫻10⫺7 kg/m
density where

冕 冕
Base mass of AFM m 0.001 kg
AFM Tip mass me 3.0⫻10⫺10 kg L L
Micro-cantilever Viscous damping B 0.1 kg/ms A i⫽ ␳ ␾ i 共 x 兲 dx⫹m e ␾ i 共 L 兲 , N i⫽ ␳ ␾ i2 共 x 兲 dx⫹m e ␾ i2 共 L 兲
Micro-cantilever Structural C 0.01 kg/s 0 0
damping
Hamacker constant 关10兴
Hamacker constant 关10兴
H1
H2
10⫺19
10⫺76
J
J Ci j⫽ 冕0
L
␾ i 共 x 兲关 B ␾ j 共 x 兲 ⫹C ␾ ⬘j 共 x 兲兴 dx, S i ⫽EI 冕 0
L
␾ i⬙ 2 共 x 兲 dx

D 1i ⫽
8
3L 冕 0
L
␾ i 共 x 兲 dx,
equations. In order to do so, the boundary conditions 共11兲 need to

冕 冕
be homogenized. For this, we define a new variable z(x,t) L L
D 2i ⫽⫺B g 共 x 兲 ␾ i 共 x 兲 dx⫺C g ⬘ 共 x 兲 ␾ i 共 x 兲 dx (33)
u 共 x,t 兲 ⫽z 共 x,t 兲 ⫹ f ID 共 t 兲 g 共 x 兲 (26) 0 0

冕 冕
where g(x), 0⭐x⭐L, is a geometrical function to be found later. L L
Substituting u(x,t) from 共26兲 into the Eqs. in 共11兲, we can get the D 3i ⫽⫺ ␳ g 共 x 兲 ␾ i 共 x 兲 dx, D 4 ⫽⫺ ␳ g 共 x 兲 dx
following conditions on g(x) in order to arrive at homogenous 0 0
boundary equations in terms of new variable z(x,t)
g 共 0 兲 ⫽0, g ⬘ 共 0 兲 ⫽0 ␺ ⫽m⫹m e ⫹ ␳ L

g ⬙ 共 L 兲 ⫽0, g 共 L 兲 ⫽0 (27) In deriving the Eqs. of motion 共32兲, the following orthogonality
conditions for the mode shapes were utilized 关30兴
g ⵮ 共 L 兲 ⫽⫺1/EI
Then, g(x) can be chosen as ␳ 冕 L

0
␾ i 共 x 兲 ␾ j 共 x 兲 dx⫹m e ␾ i 共 L 兲 ␾ j 共 L 兲 ⫽N i ␦ i j
(34)
1
g 共 x 兲 ⫽⫺
18EIL
兵 2x 4 ⫺5Lx 3 ⫹3L 2 x 2 其

which satisfies conditions 共27兲. Substituting 共26兲 into the equa-


(28)
EI 冕 L

0
␾ i⬙ 共 x 兲 ␾ ⬙j 共 x 兲 dx⫽S i ␦ i j

tions of motion and the original boundary conditions, we can


obtain the following equations and homogenized boundary where ␦ i j is the Kronecker delta. The truncated n-mode descrip-
conditions tion of the beam Eqs. 共32兲 can now be represented in the follow-
ing matrix form
␳ 共 z tt 共 x,t 兲 ⫹d̈ 共 t 兲兲 ⫹Bz t 共 x,t 兲 ⫹Cz xt 共 x,t 兲 ⫹EIz xxxx 共 x,t 兲 ⫽
¨ ⫹C⌬
M⌬ ˙ ⫹K⌬⫽Fu (35)
⫺ 兵 ␳ g 共 x 兲 f̈ ID 共 t 兲 ⫹Bg 共 x 兲 ḟ ID 共 t 兲 ⫹Cg ⬘ 共 x 兲 ḟ ID 共 t 兲

冋 册
where
⫹EIg ⵳ 共 x 兲 f ID 共 t 兲 其
(29)

冕 L ␺ A1 A2 ¯ An
共 m⫹ ␳ L⫹m e 兲 d̈ 共 t 兲 ⫹ ␳ z tt 共 x,t 兲 dx⫹m e z tt 共 L,t 兲 A1 N1 0 ¯ 0
0
¯
冕 L
M⫽ A 2 0 N2 0 ,
⫽ f 共 t 兲 ⫹ f ID 共 t 兲 ⫺ f̈ ID 共 t 兲 ␳ g 共 x 兲 dx ] ] ]  ]
0

冋 册
An 0 0 ¯ Nn
z 共 0,t 兲 ⫽z x 共 0,t 兲 ⫽z xx 共 L,t 兲 ⫽0
(30) 0 0 0 ¯ 0
m e 关 d̈ 共 t 兲 ⫹z tt 共 L,t 兲兴 ⫽EIz xxx 共 L,t 兲
0 C 11 C 12 ¯ C 1n
Now, we can use AMM for system 共29兲 and the homogeneous ¯
C⫽ 0 C 21 C 22 C 2n
boundary conditions 共30兲, i.e., z(x,t) can be expanded as
] ] ]  ]

兺 ␾ 共 x 兲q 共 t 兲

冋 册
z 共 x,t 兲 ⫽ (31) 0 C n1 C n2 ¯ C nn
i i (36)
i⫽1
0 0 0 ¯ 0
where ␾ i (x)’s are the mode shapes of a cantilever beam with a tip d共 t 兲

冦 冧
mass, and q i (t)’s are beam generalized coordinates. Using Eqs. 0 S1 0 ¯ 0 q 1共 t 兲
共29兲 and 共30兲, the discretized model of the system is governed by K⫽ 0 0 S2 ¯ 0 , ⌬⫽ q 2共 t 兲 ,

] ] ]  ] ]
␺ d̈ 共 t 兲 ⫹ 兺 A q̈ 共 t 兲 ⫽ f 共 t 兲 ⫹ f
j⫽1
j j ID 共 t 兲 ⫹D 4 f̈ ID 共 t 兲 (32a)
0 0 0 ¯ Sn
q n共 t 兲

332 Õ Vol. 126, JUNE 2004 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 7 Simulation results for the AFM tip displacement for
distributed-parameters model
Fig. 6 Simulation results for the AFM base motion for
distributed-parameters model

冋 册
performed for the distributed-parameters and lumped-parameters
models. For the distributed-parameters model, using the method

再 冎
1 1 0 D4 explained in Section 3.2, the estimated interaction force can be
0 D 11 D 21 D 31 f 共t兲 calculated. Utilizing 共16兲, the boundary condition at the tip of the
f I共 t 兲 AFM in 共11兲 can be written as
F⫽ 0 D 12 D 22 D 32 , u⫽
ḟ I 共 t 兲 m e w tt 共 L,t 兲 ⫺EIw xxx 共 L,t 兲 ⫽ f ID 共 t 兲 (40)
] ] ] ] f̈ I 共 t 兲
Since we use numerical approximation to reveal the interaction
0 D 1n D 2n D 3n force, the boundary condition 共40兲 may not be exactly satisfied.
Eq. 共35兲 can be expressed in the following state-space form However, we can compare the left-hand sides of Eqs. 共21兲 and
共40兲 in the simulations. Using Eqs. 共21兲 and 共40兲, the interaction
Ẋ⫽⌳X⫹Bu (37) force can be identified utilizing the proposed controller 共see Fig.
where 8兲. The estimator gains are taken to be ␬ 0 ⫽0.01 and ␣⫽0.0001

冋 册 再冎
for the presented results.
⌳⫽
0
⫺1
⫺M K ⫺M
I
⫺1
C
, B⫽ 再 0
M⫺1 F
, 冎 ⌬
X⫽ ˙

For the case of lumped-parameters model assumption, we can
use the equation of motion to identify the interaction force. As-
(38)
to be solved using Matlab software programming.
The effect of application of control force 共17兲 to the base of the
microcantilever can now be simulated. In order to simulate the
surface scanning by the tip of AFM, it is assumed that the distance
between base frame coordinate to the sample changes as z 0 (t)
⫽0.5⫹0.1 sin(10000t) 共␮m兲. Considering the auxiliary control
input f f (t)⫽0.05 sin(105 t) 共N兲 and the control gains k r ⫽20, k p
⫽90, and ␤ 0 ⫽0.01 for the non-contact mode, the base and tip
displacements of the AFM are demonstrated in Figs. 6 and 7,
respectively.
5.2 Lumped-Parameters Model Simulations. To compare
the results of the distributed-parameters and lumped-parameters
models, the equivalent mass, damping and spring coefficients in
the lumped-parameters configuration are taken to be
m eq ⫽m e ⫹ ␳ L/3

b eq ⫽
B
␾ 21 共 L 兲
冕 0
L
␾ 21 共 x 兲 dx (39)

3EI
k eq ⫽
L3
The equivalent damping coefficient is calculated such that the
dissipated energies for the first mode of the distributed-parameters
and lumped-parameters models become equal. Fig. 8 The interaction force identification „in †N‡… for
distributed-parameters model; „a… the left hand side of „21…, „b…
5.3 Identification of Interaction Force. In order to show the left hand side of „40… and „c… the difference between „a…
the effect of the proposed force estimator, the simulations are and „b…

Journal of Dynamic Systems, Measurement, and Control JUNE 2004, Vol. 126 Õ 333

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 2 Normalized system parameters for numerical
simulations

Properties Symbol Value Unit


Beam rigidity EI 10 Nm2
Beam length L 1 m
Beam linear density ␳ 0.1 kg/m
Base mass m 1 kg
Tip mass me 0.1 kg
Viscous damping coefficient B 0.1 kg/ms
Structural damping coefficient C 0.01 kg/s
Hamacker constant H1 10⫺7 J
Hamacker constant H2 10⫺28 J

5.4 Improved Interaction Force Measurement. To show


the importance of measuring the shear force at the tip of the beam,
each term on the left hand side of 共40兲 is plotted in Fig. 10. It is
obvious that the shear force, EIw xxx (L,t), can not be ignored
when compared to the other term in the interaction force Eq. 共40兲.

5.5 Feasibility Study of Sample-Tip Distance Estimation.


An attempt is now taken to utilize the identified interaction force
in calculating the distance between sample and AFM tip 共i.e.,
Z 0 (t)⫽z 0 (t)⫺u(L,t)⫺d(t)). For this, the normalized system pa-
rameters listed in Table 2 are utilized. These parameters are cho-
Fig. 9 The interaction force identification „in †N‡…; „a… lumped- sen to facilitate the numerical simulations using our current com-
parameters model, „b… the left hand side of „21…, and „c… the
putational capabilities. The challenges will lie in the utilization of
difference between „a… and „b…
the actual system parameters for which computational efforts be-
come more timely and complicated due to the small parameters’
values and the need for precision.
suming that the base and tip displacements are measurable, we After substituting the Galerkin relation of 共31兲 into 共30兲, we
can find the interaction force using Eq. 共1兲 as follows note that the boundary conditions in 共30兲 can be satisfied by prop-
erly selecting the geometrical function g(x). It can be shown that
F IL 共 t 兲 ⫽⫺m e ẍ⫺b 共 ẋ⫺ḋ 兲 ⫺k 共 x⫺d 兲 (41) a 6th order polynomial form for g(x) will satisfy these conditions
as well as geometrical constraints given by 共27兲. Similar to the
In order to compare the estimated interaction force from the previous case, the controller at the base is selected as a linear
lumped and distributed-parameters models, the base and tip dis- controller plus a sinusoidal offset as follows 共i.e., the AFM system
placement of AFM calculated from solving the distributed- is operated in the non-contact mode兲
parameters model is used in Eq. 共41兲. The result is depicted in Fig.
9 along with the one from distributed-parameters model and their f 共 t 兲 ⫽⫺ 共 m ␤ 0 ⫺BL⫺C 兲 w t 共 0,t 兲 ⫺k r r 0 共 t 兲 ⫺k p w 共 0,t 兲
differences. Figures 8 and 9 demonstrate that the error of using
lumped-parameters model is comparable to the value of interac- ⫹0.1 sin共 10t 兲 (42)
tion force itself calculated from lumped-parameters model 共notice with the interaction force estimator being designed as given by
the errors scale in Figs. 8共c兲 and 9共c兲兲. Eqs. 共19兲 through 共21兲. Considering the number of modes in

Fig. 10 Comparison between the shear force value and the Fig. 11 Simulation results for actual and estimated normalized
other terms in interaction force Eq. „40… interaction forces

334 Õ Vol. 126, JUNE 2004 Transactions of the ASME

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Measurement by Atomic Force Microscope,’’ JSME International Journal, Se-
ries A: Mechanics and Material Engineering, 39共3兲, pp. 456 – 462.
关4兴 Yamamoto, A., Watanabe, A., Tsubakino, H., and Fukumoto, S., 2000, ‘‘AFM
Observations of Microstructures of Deposited Magnesium on Magnesium Al-
loys,’’ Materials Science Forum, 350, pp. 241–246.
关5兴 Gahlin, R., and Jacobson, S., 1998, ‘‘Novel Method to Map and Quantify Wear
on a Micro-Scale,’’ Wear, 222共2兲, pp. 93–102.
关6兴 Miyahara, K., Nagashima, N., Ohmura, T., and Matsuoka, S., 1999, ‘‘Evalua-
tion of Mechanical Properties in Nanometer Scale Using AFM-Based Nanoin-
dentation Tester,’’ Nanostruct. Mater., 12共5兲, pp. 1049–1052.
关7兴 Sundararajan, S., and Bhushan, B., 2001, ‘‘Development of a Continuous Mi-
croscratch Technique in an Atomic Force Microscope and Its Application to
Study Scratch Resistance of Ultrathin Hard Amorphous Carbon Coatings,’’ J.
Mater. Res., 16共2兲, pp. 437– 445.
关8兴 Chen, W., Ahmed, H., and Nakazoto, K., 1995, ‘‘Coulomb Blockade at 77 K in
Nanoscale Metallic Islands in a Lateral Nanostructure,’’ Appl. Phys. Lett.,
66共24兲, p. 3383.
关9兴 Klein, D. L., McEuen, P. L., Katari, J. E. B., Roth, R., and Alivisatos, A. P.,
1996, ‘‘Approach to Electrical Studies of Single Nanocrystals,’’ Appl. Phys.
Lett., 68共18兲, p. 2574.
关10兴 Bezryadin, A., Dekker, C., and Schmid, G., 1997, ‘‘Electrostatic Trapping of
Single Conducting Nanoparticles Between Nanoelectrodes,’’ Appl. Phys. Lett.,
71共9兲, pp. 1273–1275.
关11兴 Basso, M., Giarre, L., Dahleh, M., and Mezic, I., 1998, ‘‘Numerical Analysis
of Complex Dynamics in Atomic Force Microscopes,’’ Proceedings of the
IEEE Conference on Control Applications, Trieste, Italy, pp. 1026 –1030.
Fig. 12 Simulation results for actual and estimated normalized 关12兴 Fung, R., and Huang, S., 2001, ‘‘Dynamic Modeling and Vibration Analysis of
distance between AFM tip and sample surface Atomic Force Microscope,’’ ASME J. Vibr. Acoust., 123, pp. 502–509.
关13兴 Ashhab, M., Salapaka, M., Dahleh, M., and Mezic, I., 1999, ‘‘Dynamical
Analysis and Control of Microcantilevers,’’ Automatica, 35, pp. 1663–1670.
关14兴 Hsu, S., and Fu, L. 1999, ‘‘Robust Output High-Gain Feedback Controllers for
AMM model to be n⫽3, the actual and estimated atomic interac- the Atomic Force Microscope Under High Data Sampling Rate,’’ Proceedings
tion forces are both depicted in Fig. 11. As illustrated, the esti- of the IEEE International Conference on Control Applications, Kohala Coast-
mated interaction force converges very rapidly to the actual force, Island, HI, pp. 1626 –1631.
关15兴 Banks, H., and Inman, D., 1991, ‘‘On Damping Mechanisms in Beams,’’
demonstrating the effectiveness of the proposed force estimation ASME J. Appl. Mech., 58, pp. 716 –723.
scheme. Using the estimated interaction force, the distance be- 关16兴 Dadfarnia, M., Jalili, N., Xian, B., and Dawson, D. M., 2003, ‘‘An Investiga-
tween the AFM tip and sample surface Z 0 (t) from Eq. 共8兲 can be tion of Damping Mechanisms in Translational Euler-Bernoulli Beams Using a
subsequently determined. Figure 12 presents the actual and esti- Lyapunov-Based Stability Approach,’’ Proceedings of the 2003 ASME Inter-
national Mechanical Engineering Congress and Exposition, Symposium on
mated plots of this distance. Active Vibration and Noise Control, Washington DC, November 2003.
关17兴 Fang, Y., Dawson, D., Feemster, M., and Jalili, N., 2002, ‘‘Nonlinear Control
6 Conclusion Techniques for the Atomic Force Microscope System,’’ Proceedings of 2002
International Mechanical Engineering Congress and Exposition (IMECE’02),
The non-contact AFM imaging system can significantly impact New Orleans, Louisiana.
many fabrication and manufacturing processes at molecular and 关18兴 Fang, Y., Feemster, M., Dawson, D., and Jalili, N., 2002, ‘‘Active Interaction
nanoscale levels due to its tremendous surface microscopic capa- Force Identification for Atomic Force Microscope Applications,’’ Proceedings
bilities. Moreover, non-contact imaging at low tip oscillation am- of 41st IEEE Conference on Decision Control (CDC’02), Las Vegas, Nevada.
关19兴 Slotine, J., and Li, W., 1991, Applied Nonlinear Control, Prentice Hall, Engle-
plitudes allows extremely high resolution imaging without the tip wood Cliffs, New Jersey, 1991.
damage which occurs using tapping technique. Due to its direct 关20兴 Dixon, W. E., Dawson, D. M., Zergeroglu, E., and Behal, A., 2001, Nonlinear
influence on image resolution, the distributed nature of the struc- Control of Wheeled Mobile Robots, Springer-Verlag, London Ltd.
tural flexibility in the micro-cantilever is an issue of increasing 关21兴 Kristic, M., Kanellakopoulos, I., and Kokotovic, P., 1995, Nonlinear and
Adaptive Control Design, John Wiley and Sons, Inc., New York.
concern. Since the magnitude of the interaction force lies within 关22兴 Diken, H., 2000, ‘‘Vibration Control of a Rotating Euler-Bernoulli Beam,’’ J.
the range of nano-Newtons to pica-Newtons, precise estimation of Sound Vib., 232共3兲, pp. 541–551.
the atomic force is crucial for accurate topographical imaging. In 关23兴 Luo, Z.-H., and Guo, B.-Z., 1997, ‘‘Shear Force Feedback Control of a Single-
contrast to the previously utilized lumped modeling methods, a Link Flexible Robot With a Revolute Joint,’’ IEEE Trans. Autom. Control,
42共1兲, pp. 53– 65.
general distributed-parameters base modeling approach was de- 关24兴 Kasai, S., and Matsuno, F., 2000, ‘‘Force Control of One-link Flexible Arms
veloped that revealed greater insight into the fundamental charac- Mounted on a Linear Motor With PD and Shear Force 共PDSF兲 Feedback,’’
teristics of the microcantilever-sample interaction. The proposed IECON Proceedings (Industrial Electronics Conference), 26th Annual Confer-
AFM controller interaction force estimator was then designed ence of the IEEE Electronics Society IECON 2000, Nagoya, pp. 195–200.
based on the original infinite dimensional distributed-parameters 关25兴 Luo, Z.-H., and Feng, D.-X., 1999, ‘‘Nonlinear Torque Control of a Single-
Link Flexible Robot,’’ J. Rob. Syst., 16共1兲, pp. 25–35.
system. Numerical simulations were provided to demonstrate 关26兴 Rabe, U., Turner, J., and Arnold, W., 1998, ‘‘Analysis of the High Frequency
these features and support the statements claimed here. Response of Atomic Force Microscope Cantilevers,’’ Appl. Phys. A: Mater.
Sci. Process., A66共7兲, pp. 277–282.
关27兴 Fleming, A., and Moheimani, S., 2003, ‘‘Adaptive Piezoelectric Shunt Damp-
References ing,’’ Smart Mater. Struct., 12共1兲, pp. 18 –28.
关1兴 Goeken, M., and Kempf, M., 1999, ‘‘Microstructural Properties of Superalloys 关28兴 Fleming, A., Behrens, S., and Moheimani, S., 2002, ‘‘Optimization and Imple-
Investigated by Nanoindentations in an Atomic Force Microscope,’’ Acta mentation of Multimode Piezoelectric Shunt Damping Systems,’’ IEEE/ASME
Mater., 47共3兲, pp. 1043–1052. Trans. Mechatron., 7共1兲, pp. 87–94.
关2兴 Kempf, M., Göken, M., and Vehoff, H., 1998, ‘‘Nanohardness Measurements 关29兴 Luo, Z.-H., Kitamura, N., and Guo, B.-Z., 1995, ‘‘Shear Force Feedback Con-
for Studying Local Mechanical Properties of Metals,’’ Appl. Phys. A: Mater. trol of Flexible Robot Arms,’’ IEEE Trans. Rob. Autom., 11共5兲, pp. 760–765.
Sci. Process., 66, pp. S843–S846. 关30兴 Meirovitch, L., 1997, Principles and Techniques of Vibrations, Prentice Hall,
关3兴 Nagashima, N., Matsuoka, S., and Miyahara, K., 1996, ‘‘Nanoscopic Hardness Upper Saddle River, New Jersey.

Journal of Dynamic Systems, Measurement, and Control JUNE 2004, Vol. 126 Õ 335

Downloaded From: http://dynamicsystems.asmedigitalcollection.asme.org/ on 01/27/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like