You are on page 1of 19

ATOMIC FORCE

MICROSCOPY–AFM
1. Introduction

Atomic force microscopy (AFM) is an ultrahigh-resolution microscopy technique


that has many unique features that make it an extremely powerful experimental
technique in the area of chemistry and chemical technology. Commercial atomic
force microscope instruments can easily measure the arrangement of individual
atoms on an ultraclean surface, and specialized instruments have even gone
further and imaged subatomic details and the arrangement of chemical bonds
within individual molecules (1–3). Although impressive, these feats are not of
much use in many chemical applications, where atomically clean surfaces are
not available. However, AFM has a great advantage over microscopy techniques
with similarly high resolution (mainly electron microscopy) in the flexibility of
the imaging environment. For AFM, there is no need to image in a vacuum or
to coat the sample to make it conductive. AFM can be applied in air at atmo-
spheric pressure, in a vacuum, or in a liquid. AFM can be carried out at room
temperature or at elevated or cryogenic temperatures. Also, the range of samples
that can be imaged without sample treatment is vast: from soft polymers to cera-
mics and metals, even individual molecules, no matter the atomic content. The
only real limitation on samples for AFM is that because it is a surface technique,
the sample must be absorbed to or fixed down onto some sort of surface for
imaging.
Another advantage of AFM is that it is not only an imaging technique. The
uniquely fine spatial control, which AFM has in scanning a tiny probe over a sur-
face, has been used to extend the capabilities of AFM to measure many different
properties at the nanoscale, including a wide range of mechanical properties,
electrical properties, magnetic properties, molecular interactions forces, and
even optical properties. Because of the high spatial resolution of the technique,
AFM-based measurements of these properties can be carried out in well-defined
locations on a sample or in a many locations over the sample, forming a ‘‘property
map,’’ enabling the imaging of many different properties over a sample surface.
Typically, when we think of microscopes, we imagine them focusing radia-
tion onto a sample and forming an image with transmitted or reflected light.
Both optical and electron microscopes work this way, but AFM is fundamentally
different. Because AFM works in a radically different way, there are important
differences in the way we operate the instrument, and interpret data, compared
to other microscopy techniques.

2. How AFM Works

The atomic force microscope is superficially similar to an instrument that has


been around a long time: the mechanical surface profiler. In a surface profiler,
a pointed tip is dragged along a sample and is typically attached to a lever
that moves up and down as the tip encounters features on the sample surface.
The up-and-down motion of the lever is monitored (often optically) as the tip
moves over the surface. The motion of the lever is assumed to be equal to the

Kirk-Othmer Encyclopedia of Chemical Technology. Copyright John Wiley & Sons, Inc. All rights reserved.
2 ATOMIC FORCE MICROSCOPY– AFM

Fig. 1. Schematic diagrams showing principles of (a) the mechanical surface profiler and
(b) the scanning tunneling microscope.

sample topography (or shape), and thus a profile of the sample surface is built up.
The principle of operation of a surface profile is shown in Figure 1.
The motion of the surface profiler may be one dimensional (ie, line scans are
made), or two dimensional, such that a kind of imaging occurs. This means that a
surface profiler can be thought of as a kind of microscope. In fact, a surface pro-
filer is somewhat like a simplified version of an atomic force microscope because
it uses a stylus to obtain a map of surface topography. However, the resolution of
this technique is fundamentally limited because the force between the tip and
the sample is uncontrolled. Because of this lack of force control, damage to tip
and the sample can occur. This means that in surface profilers, sharp tips cannot
be used because they would be broken quickly. Another predecessor of the atomic
force microscope was the scanning tunneling microscope, the STM (also shown in
Fig. 1). Unlike the surface profiler, in the STM, the tip does not touch the sample
surface. Instead, the tip is brought close to the surface (which must be conduc-
tive), and a tunneling current is measured between the tip and sample. To mea-
sure the tunneling current, the tip must be kept close to the sample while
scanning over it without touching. Therefore, instead of just moving the tip in
one or two dimensions, it is moved in three dimensions, typically by piezo trans-
lators, which can make precise movements. A high-speed feedback circuit
between the current measured and the z-movement of the piezos keeps the tip
the same distance from the surface at all times while also rastering the tip
over the sample surface in the x and y directions. A major drawback of STM is
that it is limited to conductive samples and works best in vacuum.
AFM is a generalization of the principles of STM, and because it is applic-
able to almost any sample, it overtook STM in popularity quickly and is now used
much more widely than STM. The principles of atomic force microscope operation
are illustrated in Figure 2.
The heart of the atomic force microscope is the probe. The probe in AFM is a
combination of a thin, flexible cantilever with an integrated tip at the end. Typi-
cally, the tip will have pyramidal shape and be very sharp, with a tip diameter of
5–20 nm. This tip interacts physically with the sample. In most imaging experi-
ments, there is a small repulsive force between the tip and the sample. The
motion of the probe, which depends on these forces, can be measured by several
techniques (4), but most commonly an optical lever is used. This setup is very
sensitive to the motions of the cantilever and is simple to implement, so it is
ATOMIC FORCE MICROSCOPY– AFM 3

Fig. 2. AFM principles of operation. The scanning in x and y is continuous. The laser,
probe, and photodetector serve to detect changes in height in the sample as this occurs.
The signal from the photodetector is fed back into the z scanner to maintain the probe-
sample force at a constant level.

used in almost all commercial AFM instruments. The signal from the optical
lever can have sub-Angstrom precision, and this signal is passed via a feedback
loop into the z movement transducer. The movement in most atomic force micro-
scopes is achieved using piezoelectric materials, which allows extremely precise
control of the motion of the probe relative to the sample (again, subnanometer
precision can be achieved). Whereas the x and y axis piezoelectrics scan in a repe-
titive pattern to raster the probe over the area of the sample of interest, the z axis
piezo is involved in the feedback loop to enable the fine control of tip-sample
force. The distance the z piezo has to move up and down to maintain the tip-sam-
ple force at the same set point is equivalent to the height (topography) of the
sample. A plot of this z height versus the x-y position of the probe gives us the
AFM image.
The combination of the following three factors using the feedback loop
explains the greatly improved resolution of atomic force microscopy compared
with profilometers:

1. Probe: the tip must be sharp to allow discrimination of small features. The
cantilever must be flexible (for contact mode) to allow for transduction of
small forces into movement.
2. Optical lever: greatly amplifies the movement of the probe so that small
heights can be measured and gentle forces are applied.
3. Piezoelectric translators: their high precision allows fine control of the
movement of the probe over the sample.

All of these features are illustrated in Figure 2.


4 ATOMIC FORCE MICROSCOPY– AFM

2.1. Components. Each of the components and their functions are


described subsequently, beginning with the probe.
Probe. As mentioned previously, the probes are now typically manufac-
tured with an integrated cantilever and tip, although some unusual designs
require attaching probes to bare levers. These integrated probes are now mass
produced by microelectromechanical systems technology, typically as 4-inch
wafers with approximately 400 probe substrates per wafer. The images of two
different types of probes are shown in Figure 3, along with a schematic showing
the parts of the probes. For contact modes, it is common to find single probe sub-
strates that host multiple probes of varying force constant. These can be conve-
nient for contact mode and force spectroscopy applications because it is
important to choose the most appropriate cantilever force constant for the appli-
cation. Because the sharpness of the probe tip governs the ultimate resolution
that can be achieved in AFM, the quality of the probe used is important, and
there are many designs on the market from a range of manufacturers (5).
However, in general, they are manufactured from silicon nitride (most often
for contact mode probes) or silicon (typically for oscillating modes). Several speci-
alty coatings can be obtained, which are often used in property measurement
modes (eg, a magnetic coating for magnetic measurements, conducting
coating for electrical measurements, etc). Other specialty probes include small
levers (for fast/gentle imaging and force measurement (6)), wear-resistant coat-
ings and materials, and probes with well-defined tip shapes for metrological
measurements.
Signal Transduction Mechanism. As mentioned previously, the most
common way of measuring the movement of the probe is an optical lever. This
usually consists of a small laser, which is focused onto the back of the cantilever
and is reflected on a position-sensitive photodetector (see Fig. 2). Usually, the
photodetector has four segments allowing the instrument to distinguish the
deflection of the cantilever in the vertical (parallel to the cantilever) and horizon-
tal (parallel to the sample surface) axes, although for most applications only

Fig. 3. Probes and probe layout. (a) The generalized layout showing probe substrate
(chip), cantilever, and tip. (b) Scanning electron microscopy (SEM) image of a probe
with rectangular cantilever for oscillating modes such as intermittent contact AFM.
(c) SEM image of a probe with ‘‘v’’-shaped cantilever for contact mode AFM. Often chips
for contact mode have multiple integrated probes to enable the selection of a cantilever
with appropriate spring constant.
ATOMIC FORCE MICROSCOPY– AFM 5

vertical deflection is used. Several other mechanisms are used (4) (notably inter-
ferometric measurements (7)), but the optical lever is employed in the over-
whelming number of atomic force microscope designs. The major advantages of
this design are that it is extremely sensitive, as well as cheap and simple to
implement. Drawbacks include the fact that it is not particularly compact and
that is somewhat more complex to implement in probe-scanning designs. There
can be some variation on the laser used; for example, low coherence lasers reduce
interference effects, whereas nonvisible (eg, infrared) lasers can be used to avoid
interfering with simultaneous optical microscopy.
Feedback Electronics. The signal from the photodetector must go
through a feedback loop into the z piezoelectric element. Typically, this is done
using a proportional, integral, derivative (PID) controller, which moves the z
translator such that the cantilever deflection returns to a setpoint. The setpoint
is programmed either automatically by the software or by the instrument opera-
tor. The P, I, and D terms can be altered in real time by the operator to optimize
imaging, along with other parameters including the setpoint and imaging speed.
Typically, the feedback is handled by an ‘‘electronics box’’ located between the
AFM head and a controlling PC, which also generates the signals for x and y
translators, and handles the digitization of the image signals.
Piezoelectric Translators. A wide variety of translation mechanisms for
use in AFM instruments has been investigated, but in general, the mechanisms
nearly always rely on piezoelectric ceramic materials. As mentioned previously,
this is because these devices can produce extremely precise movements. When
calibrated properly, they can also be capable of extremely accurate scanning;
however, various issues with piezoelectric materials can give rise to significant
nonlinearity in their real-world operation (8,9). Piezoelectric materials expand
or contract on the application of an electrical potential, and the materials used
typically have expansion coefficients of the order of 0.1 nm per applied volt. Thus,
it is simple to use piezoelectrics to scan an AFM probe with the precision
required to image individual atoms with the atomic force microscope. However,
some problems associated with piezoelectrics include nonlinear response, hyster-
esis, and creep. These problems lead to distorted images and difficulties in per-
forming zoom-to-feature, which is an important requirement for any microscope.
To overcome some of these problems, a variety of different piezoelectric device
geometries has been described, as well as the use of piezoelectrics in more com-
plex scanning devices, which are designed to improve linearity of scanning (10).
Ultimately, the best linearity is currently found in instruments that add a sensor
to the scanning mechanism and use the signal from the sensor to control the
scanning. This is usually done in a ‘‘closed-loop’’ configuration for the X-Y scan-
ning. More details about how this is done along with the description of some dif-
ferent sensor designs are given in Reference 4.
2.2. Topographic Scanning Modes. Topographic scanning is the most
important and common experiment in atomic force microscopy. Most AFM
images presented are actually topography images. Often they are shown in
three-dimensional or light-shaded views, but nonetheless these images are
based on topographic height maps. This is perhaps so widely used because,
despite all the useful complementary information that can be obtained using
other modes and signals, normally topographic images are the only ones
6 ATOMIC FORCE MICROSCOPY– AFM

containing three-dimensional quantifiable information. The feature heights mea-


sured in AFM are generally accurate, and therefore, from topographic images, it
is possible to make direct measurements of the sample dimensions.
Contact Mode. The first imaging mode developed for AFM was what is
now known as contact mode (11). This mode is the simplest, conceptually, to
understand. This mode is so called because it relies on the AFM probe tip
being in constant contact with the sample. In all AFM modes, there is a so-called
error signal. The error signal is effectively the signal used to minimize errors in
the z position of the probe at any time. In the case of contact mode, the error sig-
nal is the vertical deflection of the cantilever. Typically, the user will decide on a
setpoint for the deflection, which the feedback electronics maintain by moving
the z piezoelectric up and down during scanning. By maintaining the setpoint
constant, the tip-sample interaction force is kept constant. This force is repulsive
but can be maintained at a low level—it would normally be less than one nano-
newton and might be less than 100 piconewtons.
The force applied to the surface by the tip in contact mode is given by
Hooke’s law: F ¼ –k  D, where F¼ force (N), k ¼ cantilever force constant
(N/m), and D ¼ deflection distance (m). Therefore, to apply such small forces, spe-
cific contact mode probes are used; usually these are made of silicon nitride
(Si3N4) and have much smaller spring constants than probes for oscillating
modes, which are typically in the range of 0.05 to 0.1 N/m.
The fact that the probe tip is in contact with the sample has the following
important implications for contact mode AFM:

 The tip may damage or otherwise change the sample.


 The tip could be damaged.
 As well as the normal forces applied, lateral forces are applied between the
tip and the samples as scanning occurs.
 It is possible to make some mechanical measurements of the sample using
contact mode.

It is a widely held assumption that contact mode is a poor choice for soft
samples because of the normal forces mentioned previously. This is not always
true, and there are many examples where contact mode has been shown to be
suitable for imaging such samples as living cells (12–14), soft polymers (15,16),
and individual molecules (17–19). In fact, the lateral forces mentioned tend to
cause more problems. For example, imaging weekly adhered samples on a
solid substrate can present serious problems because of a ‘‘sweeping’’ action of
the probe associated with these lateral forces. Despite these issues and the
increasing use of other modes, contact mode is often still used today for specific
applications, and because of certain advantages it has as topographic mode. For
example, contact mode is the fastest scanning AFM mode, since the feedback sys-
tem does not have to sum oscillations before it can act. In normal circumstances,
contact mode can often scan at two or three times the speed of intermittent
contact mode. Furthermore, contact mode scanning has been developed with
high-speed scanners into so-called video rate AFM, allowing the user to collect
whole frames in milliseconds (19,20). Another strength of contact mode is high
ATOMIC FORCE MICROSCOPY– AFM 7

resolution. Contact mode has been successful in attaining true atomic resolution
(21) and submolecular resolution on individual molecules in liquid (17,21,22),
and it is used commonly for demonstrations of atomic lattice resolution, which
is simple to achieve (23–25).
Oscillating Probe Modes. It was realized at the start of the development
of AFM that advantages could be gained by oscillating the probe while scanning
(11). The advantages of this are twofold. First, monitoring the oscillation of a can-
tilever is a much more sensitive way of measuring the forces acting on it than
measuring the deflection directly. This allows the implementation of so-called
‘‘noncontact’’ techniques, where feedback can be maintained without the probe
ever coming into repulsive contact with the sample surface. The second reason
is that these modes effectively eliminate the lateral forces that exist in contact
mode AFM. In this way, it can be considerably easier to image weakly adhered
samples and to reduce tip and/or sample damage during scanning. Because of the
simplicity of operation with a wide variety of samples, these modes have become
popular and are now undoubtedly applied much more often than contact mode
for topographic imaging applications.
Oscillating probe methods use an additional piezoelectric device (sometimes
called a ‘‘dither piezo’’), close to the probe substrate, to oscillate the probe verti-
cally, usually close to its natural resonant frequency. The resonant frequencies of
commercial probes for oscillating modes typically vary between approximately 60
and 400 kHz. In most instruments, the oscillation is monitored by feeding the
signal from the photodetector into a lock-in amplifier, which measures the ampli-
tude and phase of oscillation. When the probe approaches the sample, both the
phase and amplitude of oscillation will be affected. In most instruments, the
amplitude is used as the error signal in oscillating modes. The amplitude will
decrease as the tip starts to come into intermittent contact with the sample sur-
face during oscillation. The feedback electronics would then maintain this
decrease in amplitude at the preselected setpoint. The same can be done using
the phase of oscillation as the error signal. If the probe does not actually touch
the sample, then interaction usually occurs by long-range forces damping oscilla-
tion and reducing the frequency of operation. In this case, the amplitude or phase
at a fixed frequency will typically also be altered, but it is also possible to monitor
change in frequency more directly. However, this requires additional electronics
for most instruments (26,27). Intermittent-contact AFM, which is named differ-
ently by each manufacturer of AFM instruments (28) but that typically involves
large oscillations (10–100 nm),and amplitude as the error signal, is probably the
most widely used AFM imaging mode today. It is applied widely, especially in
ambient environments, and it can be used to image almost any sample. It can
also be applied in liquid, although some complication in selecting the most appro-
priate resonant frequency can occur (29). It gives medium resolution, being
typically incapable of resolution less than 1 nm. A common application of inter-
mittent contact AFM is so-called phase imaging, which means using the ampli-
tude as the error signal and recording the phase of oscillation, which can have
sensitivity to material properties such as viscoelasticity (30–32). A related oscil-
lating mode is carried out by using direct frequency detection and very small
oscillations (1–10 nm). Under these circumstances, the technique is usually
called noncontact AFM. Noncontact AFM is often applied in vacuum, where it
8 ATOMIC FORCE MICROSCOPY– AFM

has considerable advantages and can give true atomic resolution (33,34). Non-
contact AFM can also be applied in liquids, and although it is used less often
than intermittent contact AFM, it can give extremely high-resolution imaging
in liquid (35).
2.3. Nontopographic Modes. One of the major strengths of AFM is
that there are many ‘‘secondary modes’’ that measure properties other than topo-
graphy. These modes, like the topographic modes discussed previously but unlike
force spectroscopy (discussed in the next section), are imaging modes, but the sig-
nals they generate give more information than the basic topography of standard
AFM. Many of these modes can be carried out with a standard AFM instrument,
only requiring the mode to be enabled in software or the use of special probes.
Lateral Force Microscopy. Lateral force microscopy (LFM) is a technique
that is extremely simple to carry out, requires no special equipment, and can give
material contrast on many samples. Lateral force microscopy works by measur-
ing the lateral bending (twisting) of the cantilever as it scans the sample. This is
done by comparing the segments on the left and right of the photodetector shown
on Figure 2. Twisting of the cantilever always occurs as the sample is scanned,
and it shows changes in regions of changing slope but is also sensitive to the fric-
tion force between the probe tip and the sample. For this reason, LFM is some-
times called friction force microscopy (FFM). Unfortunately, the topographic
sensitivity of the LFM signal can often complicate the interpretation of possible
frictional differences in regions of the samples, unless the sample is flat. How-
ever, lateral force microscopy has been applied successfully to measure friction
on organic monolayers (36), friction between two monolayers (37), and discrimi-
nation of phases in polymer blends (38). It is used often to generate contrast at
the atomic scale (39). With care, the resulting signal can be calibrated to give
important frictional measurements of sample surfaces (40).
Mechanical Property Modes. In addition to LFM, which can measure
friction, a variety of modes are sensitive to the mechanical properties of the sam-
ple. Probably the most commonly used of these is phase imaging, which as pre-
viously mentioned is carried out by imaging in amplitude-modulation-based
oscillating mode while recording the phase shift. This mode is certainly sensitive
to different materials, and it is thought that the contrast is based on adhesion
and viscoelastic properties of the material under study (31). Like many AFM
techniques, it is also sensitive to topographic changes, so care must be taken
in interpreting phase images. Other techniques that can generate images show-
ing mechanical properties include pulsed-force AFM (41) and jumping-mode
AFM (42). Both of these techniques involve ramping up the z position of the
probe while scanning but at a rate lower than the resonant frequency. By mea-
suring the slope of the cantilever deflection versus distance, a parameter related
to sample stiffness is obtained, which can be displayed in a separate channel to
the topography image. Similarly, the ‘‘pull-off height’’ can be plotted to show
tip-sample adhesion properties. Both of these parameters are explained in the
section on force spectroscopy because they are more often determined by force
spectroscopic (ie, one-dimensional instead of three-dimensional) experiments.
Neither jumping mode nor pulsed force mode is widely available because they
require specific instruments or additional hardware to implement, which
explains why they are used less often than phase imaging.
ATOMIC FORCE MICROSCOPY– AFM 9

Magnetic Force Microscopy. Magnetic force microscopy (MFM) is a gen-


eral term used for AFM-based methods to measure magnetic properties of the
sample. In fact, several different techniques are used to measure magnetic prop-
erties, but most use the atomic force microscope to measure the oscillation of a
magnetically sensitive probe when it is far (5–100 nm) from the sample surface.
Thus, they can measure the interaction of the probe with magnetic fields from
the samples. Usually, MFM probes are made by coating normal silicon probes
with a thin magnetic coating (eg, cobalt, cobalt-nickel, or cobalt-chromium
(43)). The magnetic coating means that the oscillation of the probe will change
when in a magnetic field. However, when the probe tip is touching the sample,
the short-range tip-sample forces (such as van de Waals forces) will obscure the
magnetic forces (which are much weaker). Fortunately, because magnetic forces
can be measured at a distance, it is possible to lift the probe from the surface and
still measure magnetic forces while removing these short-range forces. To make
accurate measurements, the probe should be at the same distance from the sam-
ple throughout the image. Several techniques are used to do this, which are
reviewed in Reference 44, but in most commercial implementations, the so-called
lifting method (45,46) is used. This method is illustrated schematically in
Figure 4.
To make MFM measurements with this lifting mode, an oscillating mode is
used, which is typically intermittent contact mode AFM. For each line of the
image, two scans are made. In the first line, the topography is measured as
usual. The probe is then lifted a user-defined distance above the surface (typi-
cally in the range of 5 to 50 nm). The second line will then be measured, but
the topography measured in the first line is added to the height of the probe
as it scans along the line. In this way, the instrument attempts to keep the
probe at exactly the same distance from the sample surface at each point in
the MFM image. Typically at the lifted height, the cantilever oscillation phase
is recorded, as this is usually more sensitive than the amplitude. MFM is used
widely to generate images of magnetic fields associated with small domains and

Fig. 4. Illustration of lifting modes used in MFM and EFM. Typically, a first scan (left)
measures sample topography, which is then used to enable scanning at a set height above
the surface (right). This second scan records the magnetic (or electrostatic) field informa-
tion, free of interference from short-range forces.
10 ATOMIC FORCE MICROSCOPY– AFM

is particularly of use in the development of magnetic recording technology (47).


It can also show magnetic fields associated with individual magnetic nanoparti-
cles (48,49). However, interpretation of MFM signals is complicated by the
unknown nature of the probe, and it is limited to measuring fairly intense fields
(49,50).
Kelvin Probe Microscopy. Many electrical modes for AFM exist, which
are driven largely by the semiconductor industry and the interest in making
electrical measurements of nanoscale devices (3). The simplest of these modes
is probably electrical force microscopy (EFM). In this mode, the electrostatic
force between the probe and the sample is measured by monitoring the
phase or amplitude of an oscillating tip far from the surface. To overcome topo-
graphic artifacts, this is typically carried out in a lifting mode, with an inter-
leaved scan line following the previously measured topography as described
for MFM. This can be carried out using a standard of conducting (metal-coated)
probe.
Scanning Kelvin probe microscopy (SKPM) is a rather more sophisticated
electrical method (51). In this mode, it is not necessary to measure the topogra-
phy before each scan line because the feedback is maintained far from the surface
and is based on the electrical properties of the sample. Thus, an image of the
sample potential can be built up, and the sample electrical work function can
be determined (52).
2.4. Force Spectroscopy. Not all AFM experiments are imaging
modes. A whole range of modes are based on making measurements with the
AFM probe while it interacts with the surface vertically (ie, perpendicular to
the sample plane). The most commonly used of these modes is force spectroscopy.
In this experiment, the force of interaction between the probe tip and the
sample surface is measured directly by bringing the probe tip into contact with
the surface and then pulling it away. Because the force is related to the cantile-
ver deflection by Hooke’s law as shown above, monitoring the cantilever deflec-
tions allows us to measure the force required to separate the probe tip from the
sample. If we have molecules of interest on the tip and on the sample surface, we
can measure the force required to separate them (ie, a molecular interaction
force).
This interaction parameter is obtained typically from a so-called force–
distance curve, an idealized version of which is shown in Figure 5.
In general, each interaction measured will be different because of such fac-
tors as different numbers of molecular interactions formed, changes in molecular
orientation, and so on. Therefore, under each experimental condition, many
(typically hundreds) curves will be measured. Thus, a range of different interac-
tion forces will be recorded, which explains the use of the term force spectro-
scopy.
Force spectroscopy can be used to probe the interaction force between dif-
ferent molecules (53) or self interaction of molecules (54), and it has also been
used to measure the distribution of forces found between bare atomic force micro-
scope tips and heterogeneous samples (55). In addition, it can be used to measure
the force of interaction between colloids (56). In this latter technique, typically a
microsphere would be glued to the end of the cantilever and the interaction with
another colloid or a model surface (57) would be measured. Force spectroscopy
ATOMIC FORCE MICROSCOPY– AFM 11

Fig. 5. Simplified force-distance curve. The measurement of the curve begins with the
probe far from surface (a), where there is no tip-sample interaction force and no deflection.
At (b), the probe is attracted to the surface and ‘‘snap-in’’ occurs as the probe jumps onto
the sample. Here, the force is attractive. As movement continues, the sample applies
repulsive forces to the probe and, thus, the cantilever deflects in the opposite sense, until
movement is reversed at (c). Typically, adhesive tip-sample interactions mean that as the
probe is retracted, the tip remains on the sample until these forces are overcome, when at
(d), ‘‘pull-off’’ occurs. As shown, the slope (sample stiffness) and adhesion (tip-sample
interaction force) data can be measured from such a curve.

can be carried out in a grid pattern over a sample surface to generate a map of
interaction forces. In this mode, it is sometimes called chemical force microscopy
(CFM) (58,59).
Nanoindentation. During the acquisition of force curves, as shown in
Figure 5, there are two phases: the approach and the retract segments. By mon-
itoring what happens during the approach phase while the probe tip is in physi-
cal contact with the sample, we can carry out nanoindentation experiments. This
quantitative technique, in principle, can give useful mechanical parameters of
the sample such as spring constant or Young’s modulus (60,61). It is possible
to carry out nanoindentation experiments and then to image the resulting
indents in the sample at high resolution to gain insight into deformation
mechanisms (62). A major advantage of AFM for nanoindentation experiments
is that it is possible to use the highly precise positioning capabilities of AFM to
carry out such experiments in specific parts of the sample. For example,
researchers made nanoindentation measurements on individual microorganisms
(63) and on nanoparticles (64). Furthermore, multiple curves can be acquired in a
grid pattern to enable the mapping of the hardness of samples, for example, on
heterogeneous polymer films (16,65,66). A disadvantage of AFM-based nanoin-
dentation experiments is that only estimated values of the properties measured
can be obtained unless calibration of both cantilever spring constant and tip-
shape are made. Data analysis can be complicated also (19). The tip-shape pro-
blem is often overcome by using a tipless cantilever with a spherical colloid glued
to it in place of the more normal sharp tip.
12 ATOMIC FORCE MICROSCOPY– AFM

3. Applications of AFM in Chemical Technology

The range of applications of AFM in chemistry is vast. It has been shown to have
many uses not only in the traditional disciplines of physical, organic, inorganic,
and analytical chemistry but also in more applied and specialized areas such as
food science, surface science, nanotechnology, biochemistry, and even in chemi-
cal industry. In this article, therefore, I will not try to cover the range of applica-
tions in which AFM is used, but instead just a few examples are described to
illustrate the capabilities of AFM with direct relevance to chemists.
3.1. Electrochemical AFM. One technique of most direct relevance to
chemistry is that of electrochemical AFM. In reality, this technique makes use
of standard AFM techniques but with a specialized sample cell to allow electro-
chemical modifications and measurements to be made simultaneously with AFM
measurements. Typically, a small. sealed liquid cell is used with connections for
the electrodes to bias the sample versus a reference electrode. These can be con-
structed by the atomic force microscope user and are commonly available for pur-
chase from the major atomic force microscope manufacturers, requiring only the
addition of a potentiostat.
With such a setup, it is possible to carry out in situ imaging of electroche-
mical processes occurring on sample surfaces, typically at metal electrode sur-
face. By ramping up the applied potential during or between AFM scans, it is
possible to observe the results of reduction and oxidation processes directly at
the electrode surface. Such processes tend to give rise to small changes, so it is
useful to use AFM to observe them. Using the AFM probe (typically with a con-
ducting coating), it is also possible to carry out electrical measurements at the
nanoscale, which is referred to as scanning electrochemical AFM. An example
showing images collected during reduction of gold on an electrode surface is
shown in Figure 6.
3.2. High-Resolution Imaging. Most AFM imaging does not actually
use the high-resolution capabilities of AFM, but instead it involves collecting
images with sizes in the range of 2 to 50 mm, meaning resolution does not nor-
mally go beyond 5 nm. One reason for this is that for useful high-resolution ima-
ging, it is necessary to have a well-organized and clean sample. Another reason is
that achieving higher resolution can be difficult, and it requires great care and
persistence. In some cases, specialized AFM techniques are required, meaning
that sometimes these capabilities are beyond the scope of a typical commercial
atomic force microscope scanning in ambient conditions.
However, this is not always the case. For example, an experiment that
is carried out routinely with atomic force microscopes to demonstrate high-
resolution imaging is atomic-lattice imaging. In this type of imaging experiment,
the atomic force microscope enables imaging of the repeating pattern of an
atomic lattice of a well-organized material. The most common material for this
experiment is muscovite mica because it is extremely simple to prepare and
has a rather large lattice periodicity, making it relatively simple to measure.
An example image of this sample showing the atomic lattice is shown in
Figure 7. In addition to mica, other samples that have been imaged in this
way include the Au (111) surface, salt crystals, highly ordered pyrolytic graphite,
and many others (21). These experiments can be carried out in ambient
ATOMIC FORCE MICROSCOPY– AFM 13

Fig. 6. Electrochemical AFM example. Images showing the morphology of a CdTe film
during electrochemical deposition of Au, at various times of applying a potential of
0.35 V. Electrochemical AFM allows the user to follow the progression of electrochemical
processes such as this in situ. Reproduced with permission from Reference 67.

conditions or in liquid at room temperature using contact mode. The only special
care that is required is the use of a sharp tip, removal of all sources of noise,
thermal equilibrium, and typically, extremely fast scanning to overcome residual
thermal expansion. It is important to remember that these images do not in
fact show true atomic resolution but instead show the average arrangement of
atoms on a surface. Thus, this ‘‘atomic lattice resolution’’ AFM imaging, as we

Fig. 7. Examples of high-resolution imaging with the AFM. (a) Example of lattice resolu-
tion on mica measured in liquid in contact mode. Inset shows the FFT pattern. Repro-
duced with permission from Reference 68. (b) True atomic resolution image of (Si, Sn,
and Pb alloy) measured using noncontact AFM in vacuum. Reproduced with permission
from Reference 69.
14 ATOMIC FORCE MICROSCOPY– AFM

shall refer to it here, does not allow imaging of individual adatoms or of atomic
vacancies.
However, true atomic resolution can be achieved with AFM, although it is
usually achieved using so-called FM detection (which allows the precise
control of the frequency of oscillation in an oscillating mode) (33) and in most
cases operation in ultra high vacuum and sometimes at cryogenic temperatures
(69–71). These conditions enable the imaging of atomic vacancies, dislocations,
stacking faults, submonolayers, and so on (2,21). Furthermore, this technique
can even reveal subatomic detail, for example, the spectacular imaging of mole-
cular orbitals (72). Some recent advances allowed atomic resolution imaging in
liquids as well (35). In addition to cases where atomic resolution can be obtained,
submolecular information is obtainable under more relevant conditions (eg, with
hydrated samples) in the case of large molecules such as proteins, protein
complexes (22,73,74), and polymers (75,76).
3.3. Molecular Interactions. One of the most important abilities of
AFM is that it can sensitively measure tip-sample interaction force. The interac-
tion between the material on the AFM probe tip and the material on the sample
surface can give useful information about the nature of the sample. Further-
more, by attaching a species of interest to the AFM tip, the force of interaction
it has with some other species on the sample can be measured. This could be, for
example, molecule–molecule interactions or colloid–colloid interaction forces, or
even the force of interaction between an animal cell and a biomaterial. Because
the atomic force microscope is so sensitive, in principle, even an interaction
between single pairs of molecules can be detected. This technique, which as men-
tioned previously is referred to as force spectroscopy, has a huge range of poten-
tial applications, from surface science, through chemistry to biology. For
chemistry, however, probably the most interesting applications are those of sin-
gle-molecule interaction studies. These studies can be used also in a mapping-
type configuration, for example, to use molecular interactions to locate single
molecules on the surface of living cells (77,78). Probably the most common use
for this type of molecule–molecule interaction is in biological systems such as lec-
tin–carbohydrate interactions or antibody–antigen binding, but small molecules
can also be probed. For example, researchers studied hydrophobic interactions
between hexadecane molecules (34,79) and carbohydrate-carbohydrate interac-
tions (54) with this technique. Typically for this application, the molecule to
interact will be bound to the tip and possibly to a flat substrate (for self-interac-
tions), often being coupled by a flexible linker to increase the chance of the
molecules binding in appropriate configurations.
3.4. Nanolithography. In addition to imaging, the ability of AFM to
control finely the position of the probe close to a surface and to control tip-sample
force can be used to modify samples. Since the first experiments in AFM, it was
noticed that the sample could be changed by interaction with the tip. In the inter-
vening years, many techniques have been developed to modify and construct
nanostructures with the atomic force microscope. Some are unsophisticated tech-
niques, such as nanoscratching, which is a simple way to created patterns in soft
materials, and assembly of nanostructures. The latter technique involves using
the AFM probe to move nanoparticles, nanotubes, and so on around on the
surface, and it can be used for applications such as wire building or moving
ATOMIC FORCE MICROSCOPY– AFM 15

nanotubes across electrodes for electrical measurements (80,81). However, when


we use the terms ‘‘nanolithography,’’ typically we are discussing somewhat more
sophisticated technique that give a greater level of control over the structures
formed. The two most common forms of nanolithography are electrical nanolitho-
graphy (often called nano-oxidation) (82) and dip-pen nanolithography (DPN)
(83). The former technique is applied usually by scanning of a sample surface
in contact or oscillating mode, while applying a bias to the tip versus the sample.
This can result in oxidation occurring at the sample surface. Most commonly
these nano-oxidation techniques are carried out on silicon surfaces, creating
silicon oxide features, although other materials have been used as well (84).
Nanoscale local oxidation is simple to control because the bias can easily be
switched on and off at will, and under certain circumstances it can create fea-
tures with line widths as small as 2 nm (85), which is currently an order of
magnitude smaller than can be achieved via optical lithography. Postprocessing
steps can be used to modify small oxide patterns with other materials (35).
The second of the techniques mentioned, DPN, is a considerably more flex-
ible technique, although it is one that can require care to apply effectively. DPN
operates in an analogous way to an old-fashioned ink pen, in that the AFM probe
tip is dipped into a solution of the ‘‘ink’’ (ie, the material to be deposited), which is
then ‘‘written’’ onto the sample by the scanning motion. This technique has been
shown to be remarkably capable of nanoscale modification with a wide range of
materials including polymers (85,86), organic molecules (87), proteins (88), and
nanoparticles (sol-based ink (89)). Some patterns made by both nano-oxidation
and DPN are illustrated in Figure 8.
3.5. AFM Characterization of Heterogeneity in Polymer Films. Poly-
mer films are ideal samples for AFM because often they require absolutely no
sample preparation at all and are exceptionally simple to image. The only possible
sample preparation steps that are required in some cases are cutting to size
if the sample is large or cleaning if it is dirty. Depending on the application,
texture parameters such as roughness can be extremely useful and are simple
to extract from AFM data. Apart from morphology, one of the most common

Fig. 8. Examples of nanolithographic patterning. (a) Patterns made by nanoxidation and


later modified to adhere single apoferritin molecules (bright dots) only on the modified
regions. Reproduced from Reference 90 with kind permission of the authors. Copyright
Wiley-VCH Verlag GmbH & Co. KGaA. Reproduced with permission. (b) A bitmap image
used as the input for a DPN routine. (c) AFM (lateral force) image of the resulting surface
patterns, which are made up of organic molecules ‘‘written’’ onto a gold surface.
16 ATOMIC FORCE MICROSCOPY– AFM

requirements with heterogeneous polymer films is visualization of the different


components of the films. Because nearly all polymer films are heterogeneous to
some degree, this capability is useful. Although some polymer blends, for exam-
ple, give rise to domain with features with different heights, in many cases this is
not so, and the different components of the films are well mixed, meaning that
topographic imaging alone is not useful to identify and quantify different compo-
nents. Fortunately, several techniques can in principle give rise to differing con-
trast based on features within a polymer film. These include lateral force
microscopy, jumping and pulsed force modes, nanoindentation mapping, and
phase mode imaging. Of these, perhaps the most useful and readily available
modes for polymer surface characterization are LFM and phase mode because
they are possible to achieve with all AFM instruments and produce contrast
while scanning at normal AFM imaging speeds. To give true frictional informa-
tion, the LFM image should be obtained by recording images scanned in both
the forward and reverse directions, and the two images subtracted from each
other to give the difference (44). By this method, contrast is often observed and
can be related to the frictional properties of the bulk material to attempt to iden-
tify the phases. This technique is simple to apply and has been applied to the
study of polymer blends (38,91), filled polymers (92), copolymers (93), and so on.
An example of frictional contrast measured on a filled polymer surface is shown
Figure 9.
Probably the most commonly used mode for this application is phase ima-
ging. One reason for this is the widespread usage of intermittent contact AFM
modes, which typically give direct access to phase imaging. However, another
reason is that the phase signal in intermittent-contact AFM is extremely sensi-
tive to changes in the viscoelastic properties of the materials studies, and thus, it
very frequently gives high contrast on polymer samples, even where there is no
discernable height difference between the components. Indeed, the most common
‘‘reference sample’’ for phase mode imaging is a copolymer sample with different
phases, which give a high contrast, enabling the user to check that phase

Fig. 9. Friction and phase imaging of examples. (a) Friction image (difference of left to
right and right to left scans) on a filled polymer films showing low friction on additive
articles. Adapted with permission from Reference 92. (b) Topography image of crystalliz-
ing polymer with spherulites, which give low contrast in height. (c) Phase image of the
same area, showing clear contrast on the spherulites. Reproduced with permission from
Reference 94.
ATOMIC FORCE MICROSCOPY– AFM 17

imaging is working well (95). Because it is simple to apply and sensitive, phase
imaging has been applied to a wide range of polymer samples, including block
copolymers (96,97), imaging spherulites in crystallizing polymers (94), polymer
blends (98), and more. However, in general it is difficult to identify features
using phase imaging without some prior information to help, such as the propor-
tions of the various components, and so on (98). Even with prior knowledge of the
response of the bulk materials in phase imaging, it can be problematic to identify
the material using phase imaging definitely because the phase signal is highly
dependent on experimental setup and phase signal reversal can even occur
under certain circumstances (99).

4. Conclusion

In summary, atomic force microscopy is an extremely useful technique in all


areas of chemistry, combining extremely high-resolution imaging with the
measurement of a wide range of physical properties in a localized manner. In
comparison with other high-resolution microscopy techniques, AFM stands out
because of its ability to measure in almost any environment, giving rise to a flex-
ible methodology and the possibility to make extremely relevant measurements,
even during chemical reaction processes.

BIBLIOGRAPHY

‘‘Atomic Force Microscopy-AFM’’ in ECT 5th ed., Vol. 5, pp. 319–341, by Charles Anthony
Peterson, Intel Corporation and Dror Sarid, Optical Sciences Center, University of
Arizona.

1. F. J. Giessibl and co-workers, Science 289, 422–425 (2000).


2. F. J. Giessibl and C. F. Quate, Phys. Today 59, 44–50 (2006).
3. R. A. Oliver, Rep. Progr. Phys. 71, 076501 (2008).
4. P. Eaton and P. West, Atomic Force Microscopy, Oxford University Press, Oxford,
U.K., 2010, Chapt. 2, pp. 9–48.
5. G. Kaupp and co-workers, Workshop on Development and Industrial Application of
Scanning Probe Microscopes (SXM-1), Elsevier Science SA, Lausanne, Switzerland,
1994, pp. 205–211.
6. M. B. Viani and co-workers, Rev. Sci. Instrum. 70, 4300 (1999).
7. L. A. Bottomley, Anal. Chem. 70, 425R–475R (1998).
8. P. Eaton and P. West, Atomic Force Microscopy, Oxford University Press, Oxford,
U.K., 2010, pp. 126–131.
9. H. Xie, M. Rakotondrabe, and S. Regnier, Rev. Sci. Instrum. 80, 046102-3 (2009).
10. C. A. J. Lin and co-workers, J. Mat. Chem. 17, 1343–1346 (2007).
11. G. Binnig, C. F. Quate, and C. Gerber, Phys. Rev. Lett. 56, 930–933 (1986).
12. M. F. Murphy and co-workers, Microsc. Res. Tech. 69, 757–765 (2006).
13. C. Le Grimellec and co-workers, Biophys. J. 75, 695–703 (1998).
14. G.-J. Lee and co-workers, Micron 41, 220–226 (2010).
15. S. H. Kim and co-workers, Biomaterials 23, 1657–1666 (2002).
16. P. Eaton and co-workers, Langmuir 18, 10011–10015 (2002).
18 ATOMIC FORCE MICROSCOPY– AFM

17. J. X. Mou and co-workers, FEBS Lett. 381, 161–164 (1996).


18. R. P. GoncSalves and co-workers, J. Struct. Biol. 149, 79–86 (2005).
19. I. Casuso and co-workers, Biophys. J. 97, 1354–1361 (2009).
20. A. D. L. Humphris, M. J. Miles, and J. K. Hobbs, Appl. Phys. Lett. 86, 034106 (2005).
21. Y. Gan, Surf. Sci. Rep. 64, 99–121 (2009).
22. D. J. Muller and co-workers, J. Struct. Biol. 119, 149–157 (1997).
23. Y. Gan, E. J. Wanless, and G. V. Franks, Surf. Sci. 601, 1064–1071 (2007).
24. C. Tromas and co-workers, Langmuir 21, 6142–6144 (2005).
25. M. Jaschke and co-workers, J. Phys. Chem. 100, 2290–2301 (1996).
26. S. Morita and co-workers, in Nanotribology and Nanomechanics, An Introduction,
Springer, New York, 2005,pp. 141–183.
27. S. Morita, R. Wiesendanger, and E. Meyer, Noncontact Atomic Force Microscopy,
Springer, New York, 2002.
28. For example, it can be known as AC-AFM, IC-AFM, vibrating mode, or tapping mode
AFM.
29. T. E. Schaffer and co-workers, J. Appl. Phys. 80, 3622–3627 (1996).
30. R. Garcia, J. Tamayo, and A. San Paulo, Surf. Interface Anal. 27, 312–316 (1999).
31. J. Tamayo and R. Garcia, Langmuir 12, 4430–4435 (1996).
32. I. Schmitz and co-workers, Applied Surf. Sci. 115, 190–198 (1997).
33. R. Garcia and R. Perez, Surf. Sci Rep. 47, 197–301 (2002).
34. Y. Sugimoto and co-workers, Appl. Surf. Sci. 241, 23–27 (2005).
35. T. Fukuma and co-workers, Appl. Phys. Lett. 87, 034101-3 (2005).
36. Q. Zhang and L. A. Archer, Langmuir 21, 5405–5413 (2005).
37. G. J. Leggett, Anal. Chim. Acta 479, 17–38 (2003).
38. P. Cyganik and co-workers, Surf. Sci. 507–510, 700–706 (2002).
39. E. Gnecco and co-workers, Phys. Rev. Lett. 84, 1172 (2000).
40. G. J. Leggett, N. J. Brewer, and K. S. L. Chonga, Phys. Chem. Chem. Phys. 7,
1107–1120 (2005).
41. H.-U. Krotil and co-workers, Surf. Interface Anal. 27, 336–340 (1999).
42. P. J. de Pablo and co-workers, Appl. Phys. Lett. 73, 3300–3302 (1998).
43. L. Abelmann and co-workers, J. Magn. Magn. Mater. 190, 135–147 (1998).
44. P. Eaton and P. West, Atomic Force Microscopy, Oxford University Press, Oxford,
U.K., 2010.
45. C. W. Lin, F.-R. F. Fan, and A. J. Bard, J. Electrochem. Soc. 134, 1038–1039 (1987).
46. R. Giles and co-workers, Appl. Phys. Lett. 63, 617–618 (1993).
47. S. Porthun, L. Abelmann, and C. Lodder, J. Magn. Magn. Mater. 182, 238–273
(1998).
48. X. Chen and co-workers, Anal. Chem. 81, 7009–7014 (2009).
49. S. Schreiber and co-workers, Small 4, 270–278 (2008).
50. C. S. Neves and co-workers, Nanotechnology 21, 305706 (2010).
51. M. Nonnenmacher, M. P. Oboyle, and H. K. Wickramasinghe, Appl. Phys. Lett. 58,
2921–2923 (1991).
52. H. O. Jacobs and co-workers, J. Appl. Phys. 84, 1168–1173 (1998).
53. H. Clausen-Schaumann and co-workers, Curr. Opin. Chem. Biol. 4, 524–530 (2000).
54. C. Tromas and co-workers, Angew. Chem., Int. Ed. 40, 3052–3055 (2001).
55. P. Eaton and co-workers, Langmuir 16, 7887–7890 (2000).
56. W. A. Ducker, T. J. Senden, and R. M. Pashley, Nature 353, 239–241 (1991).
57. W. A. Ducker, T. J. Senden, and R. M. Pashley, Langmuir 8, 1831–1836 (1992).
58. A. Noy, D. V. Vezenov, and C. M. Lieber, Annu. Rev. Mater. Res. 27, 381–421 (1997).
59. E. Dague and co-workers, Nano Lett. 7, 3026–3030 (2007).
60. C. B. Volle and co-workers, Colloids Surf. B 67, 32–40 (2008).
61. C. Gomez-Navarro, M. Burghard, and, K. Kern, Nano Lett. 8, 2045–2049 (2008).
ATOMIC FORCE MICROSCOPY– AFM 19

62. Y. Gaillard, C. Tromas, and J. Woirgard, Acta Mater. 51, 1059–1065 (2003).
63. P. Eaton and co-workers, Ultramicroscopy 108, 1128–1134 (2008).
64. H. P. Wampler and A. Ivanisevic, Micron 40, 444–448 (2009).
65. M. S. Bischel and co-workers, J. Mater. Sci. 35, 221–228 (2000).
66. M. R. Vanlandingham and co-workers, J. Adhes. 64, 31–59 (1997).
67. R. Vidu, J.-R. Ku, and P. Stroeve, J. Colloid Interface Sci. 300, 404–412 (2006).
68. Y. Kuwahara, Phys. Chem. Miner. 28, 1–8 (2001).
69. Y. Sugimoto and co-workers, Nature 446, 64–67 (2007).
70. R. Garcia, Wiley VCH Verlag GmbH & Co. KGaA, Berlin, Germany, 2010,Chapt. 1.
71. L. Gross and co-workers, Nat. Chem. 2, 821–825 (2010).
72. F. J. Giessibl and co-workers, Annalen Der Physik 10, 887–910 (2001).
73. J. X. Mou and co-workers, Biophys. J. 71, 2213–2221 (1996).
74. D. J. Muller and Y. F. Dufrene, Nat. Nanotechnol. 3, 261–269 (2008).
75. J. Kumaki, S. Sakurai, and E. Yashima, Chem. Soc. Rev. 38, 737–746 (2009).
76. N. H. Thomson, in B. Bhushan, H. Fuchs, and S. Hosaka, eds., Applied Scanning
Probe Methods, Vol. VI: Characterization, Springer-Verlag, Berlin, Germany,
2006, pp. 127–164.
77. Y. Gilbert and co-workers, Nano Lett. 7, 796–801 (2007).
78. L. A. Chtcheglova and co-workers, Biophys. J. 93, L11–L13 (2007).
79. C. Ray and co-workers, J. Phys. Chem. C 112, 18164–18172 (2008).
80. C. Ho-Yin, X. Ning, Z. Jiangbo, and L. Guangyong, 5th IEEE Conference on Nano-
technology, Vol. 2, 2005, pp. 713–716.
81. T. Junno and co-workers, Appl. Phys. Lett. 72, 548–550 (1998).
82. D. Stievenard and B. Legrand, Prog. Surf. Sci. 81, 112–140 (2006).
83. D. S. Ginger, H. Zhang, and C. A. Mirkin, Angew. Chem., Int. Ed. 43, 30–45 (2004).
84. X. N. Xie and co-workers, Mater. Sci. Eng. R 54, 1–48 (2006).
85. R. V. Martinez and co-workers, Nano Lett. 7, 1846-1850 (2007).
86. A. Noy and co-workers, Nano Lett. 2, 109–112 (2002).
87. R. D. Piner and co-workers, Science 283, 661–663 (1999).
88. K. L. Christman, V. D. Enriquez-Rios, and H. D. Maynard, Soft Matter 2, 928–939
(2006).
89. L. Fu and co-workers, Nano Lett. 3, 757–760 (2003).
90. R. V. Martinez and co-workers, Adv. Mater. 22, 588–591 (2010).
91. C. Ton-That and co-workers, Polymer 42, 1121–1129 (2001).
92. B. D. Beake, G. J. Leggett, and P. H. Shipway, Surf. Interface Anal. 27, 1084–1091
(1999).
93. S. E. Morgan, R. Misra, and P. Jones, Polymer 47, 2865–2873 (2006).
94. J. Xu and co-workers, Macromolecules 37, 4118–4123 (2004).
95. Spm refererences and standards (afmhelp.com) http://bit.ly/spm_standards.
Accessed 25th May 2011.
96. Z. W. Xu and co-workers, Ultramicroscopy 105, 72–78 (2005).
97. L. Peponi and co-workers, Macromol. Mater. Eng. 293, 568–573 (2008).
98. D. Raghavan and co-workers, J. Polym. Sci. Polym. Phys. 39, 1460–1470 (2001).
99. G. Bar and co-workers, Langmuir 13, 3807–3812 (1997).

PETER EATON
Requimte

You might also like