Вы находитесь на странице: 1из 21

spacing

Scanning Probe Techniques


Heng-Yong Nie

STM | AFM | Contact AFM | Force curve | Lateral force | Force modulation | Local
modification
Non-contact AFM | Phase imaging | Magnetic force | Surface potential | Check AFM tips
Cleaning by UVO | Conclusion | Reference & AFM manufacturers | Other info
spacing

Scanning probe microscopy (SPM) is a mechanical


probe microscopy that measures surface morphology in
real space with a resolution down to atomic resolution.
SPM was originated from the scanning tunneling
microscopy (STM), in which electrical current caused
by the tunneling of electron through the tip and sample
is used to maintain a separation between them. This
technique was a totally new one that can image atom
arrangement on a surface in real space for the first time.
It is so invaluable to science and technology related to
surface phenomena that the inventors of STM shared
the Nobel Prize in Physics 1986 with the inventor of
electron microscope.
Because STM requires that the sample surface be
conductive, atomic force microscopy (AFM) was
developed in 1986 to measure surface morphology that
are not a good conductor. AFM has since been
developed very rapidly and has found much more
applications than STM in many fields. Almost all kind
of materials can be measured by AFM. Besides surface
morphology mapping, SPM has been developed in the
past two decades to measure a variety of surface
properties, such as electrical, magnetic, and mechanical
properties. The diversity of SPM is based on the fact that the probe tip is in contact or close to
sample surface so that many interactions between the tip and sample are measurable.
Return to Top
1. STM
The principle of the STM may be simple: tunneling
of electrons between two electrodes under electric filed.
However, to develop the concept of electron tunneling
into a technology to image atomic resolution on a
surface was not simple. To measure the tunneling effect,
the distance between the two electrodes must be close to each other on an order of 1 nm.
Surface cleanness and vibration-free system are essential to the measurement of the tunneling
current. Shown here is an STM inage obtained on an HOPG substrate.

Quantum mechanics predicts an exponential dependence of tunneling current with a separate


distance between the two electrodes. An observation of this dependence between a W-tip and Pt
surface by G. Binnig, H. Rohrer, Ch. Gerber and E. Weibel in 1981 was marked as the
invention of STM. Atomic resolution measured on Si(111) 7x7 surface in 1982 might be
considered the breakthrough of STM. Since then there have appeared a huge number of papers
that carry STM images.

STM, as a research approach, has been mainly used to measure atomic resolution or
electronic structure of solid surfaces in UHV. An example of iron atoms located on Cu(111)
surface is from an IBM's research laboratory. In many research fields where the sample is
usually an insulator, AFM has been used widely.
Return to Top
2. AFM
Examples of AFM images obtained on four different samples

To obtain similar
resolution as in STM for
insulating surfaces, AFM was
invented in 1985 by G.
Binnig, C.F. Quate and Ch.
Gerber. A sharp tip (apex
radius ~20 nm) formed on a
soft cantilever is used to
probe the interaction (force)
between the tip and sample
surface, which could be
understood through the
Lennard-Jones potential
which deals with interaction
between two atoms: w(r) =
-A/r6 + B/r12, where r is the
separation of the two bodies,
A and B are interaction
constants. Then the
interaction force is F = -dw(r)/dr= -6A/r7 + 12B/r13. Following a text book, A and B are known
to be10-77Jm6 and 10-134Jm12, respectively. A calculation for the interaction force between two
atoms is shown to the right. Around a separation distance of 0.4 nm between the two atoms, a
small attractive force is seen and when the separation distance gets smaller and smaller the
repulsive force increases steeply.
For practical AFM probe tip
and the sample surface, attractive
force between them could be
much larger than what is
described here for a tow-atoms
system. This is because, at least,
the size of the tip is much larger
than an atom. Typical radius of a
commercial tip is ~10 nm. Also,
much longer-range forces could
occur in practice.

A sharp tip formed on the free


end of a cantilever is used to
probe the sample surface. The
interaction between the tip and the
surface is detected by measuring
the deflection of the cantilever
using a laser diode to radiate the
cantilever and a photodiode to
detect the reflected laser beam.
The quadrant photodiode is able to
measure both the deflection and
torsion of the cantilever. The
principle of AFM is shown in the
figure to the left. Shown in the
figure is the case where the tip scans the sample surface. The AFM operates by keeping constant
the interaction between the tip and sample surface through a feedback system that adjusts the
distance between the tip and the sample surface. Depending on the interaction between the tip
and sample surface, which is used as the feedback signal, there are two different imaging modes
described as follows

Schematic illustration of AFM principle: while scanning the tip across the sample surface (x,
y), the system adjusts the distance (z, which is thus the measure of the height of the sample
surface features) between the tip and the sample surface to maintain a constant contact force
(contact mode) or oscillation amplitude (dynamic force mode). A 3-D image is thus constructed
by the lateral dimension the tip scans and the height the system measures.
Return to Top
2.1 Contact AFM
In the contact mode AFM, the tip is mechanically contacted with sample surface, exerting a
force on the surface of the sample. This applied force can be evaluated from a force-distance
curve which is measured when the tip is brought to and then retracted from the sample surface,
as shown below. Inserts in the figure show the interaction between the tip and sample surface,
which is detected by the deflection of the cantilever. There is no interaction between the tip and
surface when the tip is far away from the surface (a ). When the tip is brought close enough to
the surface there will be an attractive force between them. Usually, the gradient of the attractive
force is much larger than the spring constant of the cantilever, so that the tip is snapped to the
surface to make a contact between the tip and surface (b). Further extending the tip results in
loading (repulsive) forces to the surface (c). This repulsive force is usually used as the feedback
parameter for the AFM system to obtain surface morphology. Forces of a couple of nN are used
in contact mode AFM. In the retracting cycle (d and e), because of the adhesion established
after the contact between the tip and surface, the tip will not detach from the surface until the
force used to pull the tip from the surface exceeds the adhesion force between them (f). This
pull-off force can serve as a measure of the adhesion force between the tip and surface.

A very soft cantilever with a spring constant of ~ 0.1 N/m is usually used in contact AFM. A
photograph of such a cantilever is shown in the optical picture below on the left. The cantilever
is so soft that it will be pulled onto the surface because the gradient (~ 10 N/m) of attractive
force between them is usually much larger than the spring constant of such soft cantilevers.
After a mechanical contact between the tip and the sample surface, there is a repulsive force
between them. This force is used as the feedback parameter (by maintaining a constant force
through adjustment of the sample height while the tip scans the surface) to obtain AFM images.
Because the tip
is mechanically
contacted with
surface in the
contact mode
AFM, many
surface properties
such as friction
force distribution
and mechanic
properties can be
measured
simultaneously
with the
topographic image.
Also, nano-
lithography on
some materials is
also available by
controlling the
applied forces in the contact mode. A working knowledge on force-distance curve is essential
for undestanding and interpreting the imaging mechanism of contact mode AFM (especcially
when things go wrong).
Return to Top
2.1.1 Force-distance curves
Force-distance curves are obtained by extending the tip to the surface to make a contact
between the tip and the sample surface followed by retracting the tip from the surface. The
original point for the distance may be defined as the mechanical contact between the tip and
surface in the extending cycle. Extending the tip beyond that point will result in load forces
applied to the surface. The slope of this load force is a measure of the Young's modulus of the
surface, possibly mixed with the spring constant of the cantilever. As a result, a cantilever
whose spring constant is comparable with the surface stiffness should be used to measure the
elasticit y information.

In the retracting cycle, because of the adhesion properties between the tip and surface, the tip
will not depart from the surface until the force used to pull the tip from the surface exceeds the
adhesion force between them. This pull-off force can be considered as a measure of the
adhesion force between the tip and surface. Adhesion force can be related to surface energies of
the tip and sample surfaces, as well as their interfacial energy. Shown above is an example of
measruing adhesion force at different regions on a BOPP film. The striped areas have higher
adhesion force than the normal surface; we will see later that this is also reflected in the friction
force images.

Click here to see adhesion force increase for UV/ozone treated polypropylene films.
Adhesion force can be related to surface energies of the tip and sample surfaces, as well as
their interfacial energy. If there were liquid-like contamination on a surface, the capillary force
should be considered. Recently, force-distance curve has been demonstrated to be able to record
the event of (a) the breaking of a single molecular bond (single molecule force spectroscopy)
and (b) the folding and unfolding of proteins by confining the molecules between the AFM tip
and a surface.

Return to Top
2.1.2 Lateral
Force Microscopy
(LFM)
Lateral force microscopy
(LFM) is based on measuring
the torsional movement of the
cantilever when the tip scans
the surface, which is illustrated
here. Lateral force detection in
AFM is usually used to image
different friction forces on a surface. The difference in the bi-directional lateral force images
corresponds to the friction force image. Friction force imaging has the ability to identify such
regions of higher hydrophilicity on the basis of increased interaction with the AFM tip. Here is
an example showing higher friction force on scratched areas on a BOPP film, which is thought
to be due to higher surface energy of the scratched area. Force-distance curves obtained on the
normal and striped areas are shown above, revealing that the friction force contrast seen is
related to the adhesion force.

Friction force can also be also used to detect chemical


functional groups on a surface with the tip modified by a
specific chemical functional base (chemical force microscopy).
For example, a silicon tip, which is hydrophilic, can be used to
d istinguish the chemical amphiphilicity of
Octadecylphosphonic acid (OPA) molecular layers. That is, film
surface terminated by the hydrophilic headgroup shows larger
friction force than that terminated by the hydropho bic tail.
Combined with height measurement, one can clearly identify
whether a molecular layer is a bilayer (terminated by the
headgroup) or a trilayer (terminated by the tail).

On the other hand, there is another effect of lateral force


imaging to reveal local topography change by an enhancement
of the torsional movement of the cantilever when the tip crosses
edges of the surface features. Click here to see an LFM image
on 15-min-UV/ozone-treated polypropylene film, where the
droplets are distinguished from the surrounding surface. This
technique has proven useful in detecting different phases on a
surface whose height range is large, which is the case for some
practical polymer samples.
Return to Top
2.1.3 Force Modulation
In addition to the topographic feature,
one can probe local elastic properties of
materials through a mechanical
interaction between the surface and tip.
This can be done by oscillating the
sample height while measuring the
response of the cantilever with lock-in
amplifier technique. Elasticity difference
on a s urface can be distinguished by
using this technique. This technique
actually measures the slope of the force-
distance curves at the repulsive force
region.

The oscillation of the sample height


may be realized by applying sinusoidal voltage from a function generator to the Z-direction of
the piezo (PZT) scanner on which the sample is fixed. A sinusoidal voltage is applied to the
piezo scanner to oscillate the sample height with a peak-to-peak amplitude of about 1 nm. The
response of the cantilever to this oscillation is detected with a lock-in amplifier and is used to
obtain images relevant to local elasticity of sample surface. The oscillation of the sample height
would not influence topographic images as far as its frequency is higher than the cutoff
frequency of the feedback loop. Therefore, both topography and elasticity distribution images
can be obtained simultaneously. An example of elasticity mapping of PS and PS/PEO blend
film coated on mica substrate using a cantilever with a spring constant of 0.75 N/m is shown in
the figure to the left.

The mechanism for the force


modulation is described using force-
distance curves (a) obtained on the
mica and the PS film and the
simultaneously obtained response (b)
of the cantilever to an oscillation of the
sample height with an amplitude of 1
nm at 5 kHz. The sprin constant of the
cantilever used was 18 N/m and the
approachoing and retracting speed for
the tip was 3 nm/s. The difference seen
for the cantilever response (b) is due to
the different slope of the force-distance (a) curves on the different materials, which is a
reflection of difference in Young’s modulus for mica (200 GPa) and PS (5GPa).

Return to Top
2.1.4 Locally modifying surface
Surface may be modified by applying large forces through the tip to the surface during an
AFM scanning. This technique has a potential application to create nanometer-scale structure on
a surface. For example, on the crystallized polyethylene oxide (PEO) thin films, both the
surface structure and elasticity were found to be modified locally by the AFM tip. A close look
at the modification of the PEO surface is shown here.

Return to Top
2.2 Dynamic force mode AFM techniques
Dynamic force (tapping or non-contact) mode AFM, in which a cantilever oscillated around
its resonant frequency is used to probe surface features, was developed initially to eliminate
surface degradation encountered in contact mode AFM, especially for soft materials. For
dynamic force mode AFM, silicon cantilevers with a spring constant of 5 ~ 40 N/m are used. A
typical 40 N/m cantilever is 125 µm long, 30 µm wide and 3.7 µm thick. Because the variation
of the oscillation amplitude is used in the feedback system, the relative change in oscillation
amplitude of the cantilever versus distance between the tip and sample surface is shown in the
figure below. The amplitude-distance curve shown in the figure was obtained on a BOPP film
surface. Interaction between the tip and the surface at different tip-sample distance is indicated
by inserts a-c. Arrows indicate the direction of the tip approaching to the sample surface.

The above figure shows that when the tip is far away from the sample surface (a), the
oscillation amplitude of the cantilever is a constant, representing a “free space” situation where
there is no interaction between the tip and the surface. The amplitude decreases when the tip
approaches close enough to the sample surface so that it “feels” attractive and/or repulsive
forces (b). The cantilever stops oscillating when the tip is brought in to mechanically contact the
surface (c). Dynamic force mode AFM works by scanning the tip across the sample surface and
adjusting the distance between the two through maintaining constant damped oscillation
amplitude of the cantilever. This adjustment of the separation between the tip and surface allows
the AFM to construct the topographic image. There are many modes measuring surface
properties based on this dynamic force mode AFM, as described in the following. Usually, a set
point at 50% of the oscillation amplitude in free space is a good start. AFM images shown
here clearly show formation of mounds on UV/ozone treated polypropylene (PP) film from the
original surface characterized by fiber-like network structure (scan area is 2 micron square and
height range is ~ 25 nm) and an increase in adhesion force. This increase in adhesion force
indicates an increase in surface energy due to the oxidation of the modified polymer films.
When the oscillating amplitude is large (say, >2 nm), the tip actually taps the surface, which
is why it is called dynamic force or tapping mode AFM. In practice of measuring larger scale
area, larger amplitude is usually used because it results in more stable imaging. If the oscillating
amplitude is very small (say, a couple of nm), then the dynamic force mode could be called
non-contact mode because at such small amplitude, the tip would not need to tap the surface to
sense the interaction between them. Operation of non-contact mode AFM is achieved in UHV
for obtaining atomic resolution through frequency-modulated techniques.
Return to Top
2.2.1 Phase Imaging
The phase shift in the oscillating cantilever is related to tip-surface interaction which is
basically material specific. Therefore, phase shift contrast in tapping mode AFM can be used to
distinguish different surface compositions on a surface (see the schematic below). There are
many surface properties that may have an effect on the phase shift contrast. They could be
difference in friction, viscoelasticity, adhesion, material, etc. Phase imaging usually gives clear
contrast on a surface if there are detectable differences in surface properties as described above.
So, the explanation of a phase shift image should be careful and usually depends on other
observations and background knowledge on the sample. It should be noted that phase imaging is
a very valuable approach for SPM researchers because they probably find and in fact are finding
some new phenomena during their searching answers and explanation to the phase shift
measurement. Applications include visualizing phase separation in polymer blends,
distinguishing different compositions on surface. Shown below is topography (left) and phase
image (right) for a surface of a tonner particle of carbon black matrix with polymer filler (scan
area is 3.5 micron square).

Return to Top
2.2.2 Magnetic Force Microscopy (MFM)
Magnetic information on a sample can be measured with a magnetized tip. A local
topographic data in each scan line is first obtained by dynamic force mode AFM. Then the tip is
lifted up in a certain distance and repeats to scan the same line. The magnetic interaction
between the tip and surface will give a change in the magnitude (or phase) of the oscillating
cantilever, which gives regional information of magnetic force distribution on a surface. The
difference of the magnetic properties can be measured in this method. This technique is useful
to image magnetic force distribution on the recorded magnetic media (data storage) and
micromagnetic structure on some magnetic materials. For more information on this topic, you
may want to visit this web page.

2.2.3 Electric Force Microscopy (EFM)


Similar to MFM, EFM uses a conductive tip to probe the difference of electric filed
gradient distribution on a surface. This technique can be used in the failure check on integrated
circuit (IC). For more information, visit this web page.
Return to Top
2.2.4 Scanning Surface Potential Microscopy (SSPM)
There has
appeared
recently a new
technique which
maps local
surface
potential
distribution
together with
topography
using an SPM,
by keeping a
certain
separation
between the
sample surface
and conductive
tip to which a
sinusoidal
voltage is
applied. The
principle of the
SSPM is shown in the figure on the right. In case of that there is a difference in potentials
between the tip and sample surfaces, an oscillating electromagnetic force appears between the
tip and sample surface at the frequency of the applied sinusoidal voltage, which makes the
cantilever oscillate. This oscillation is used as the feedback parameter for the system which tries
to stop this oscillation by applying a dc voltage to the tip so as to make the potential difference
between the tip and sample surfaces vanish. This applied dc voltage to the tip is thus equal to
the surface potential of the sample, which makes the surface potential measurable together with
the topography.
As
shown
in the
figure
to the
left,
gold
films

deposited on glass substrate were used to confirm the SSPM by making potential difference
between the two gold films through applying voltage to one of them and grounding the other.
During scanning, the voltage was changed so that different potential difference were recorded.
Metal Pd deposited on a semiconductor is an example to give contact potential between the
metal and the semiconductor surface. Here is a result showing the contact potential difference
between the Pd and semiconductor surface. Other examples of measuring surface potential
distribution are a Pd (110) surface and a thin film giving a clear distribution of surface
potential.

This technique can be used to map surface voltage distribution, which can be used to detect
defects and to measure local work function distribution.

Return to Top
3. A Simple Method to Check AFM Tip Performance Uing
a Polymer Film
3.1 BOPP film surface for tip radiu evaluation
An atomic force microscopy (AFM) image of a surface is constructed through the detection
of an interaction between the tip apex and the surface features. The interaction, whether it be a
contact force, an oscillation amplitude or others, is the feedback signal used to adjust the
proximity of the tip and the surface features. Because of this imaging mechanism, an AFM
image is, in practice, a convolution of the tip geometry and the surface features. Based on the
actual geometry, the tip apex or the surface feature, whichever is sharper, acts as the effective
probe.

In practice, there could be a large-sized contaminant on the tip apex, making sharper surface
features the effective probe. Therefore, images collected using a contaminated or damaged tip
can be dominated by the geometry of the AFM tip itself (i.e., self-imaging of the tip) if the
surface features are sharper than the tip. Interpretation of such images can easily be misleading
if the tip effect is not taken into account. To ensure that the tip is “good” enough for imaging a
surface, one needs reference samples that have known surface features, suitable for checking the
tip performance. Introduced here is a simple and effective method of evaluating tip performance
by imaging a biaxially-oriented polypropylene (BOPP) film, which is characterized by
nanometer-scale sized fibers. The BOPP film surface is appropriate for use as a reference
because a contaminated tip will not detect the fiber-like network structure. Imaging the very fine
fiber-like structure of the BOPP film surface is a good criterion for the tip performance. Many
other samples with known surface features can also be used to chracterize the geometry of
AFM probes.
Because the polymer film is soft compared to the silicon tip (Young's modulus for
polypropylene is 1-2 GPa, while for silicon it is 132-190 GPa), the polymer will not damage the
tip when the tip is pushed into the polymer. This property can be used to clean a contaminated
tip, i.e., by pushing the contaminated tip into the polymer, contaminants could be removed from
the tip apex. Another important property of the BOPP is that the polymer film is highly
hydrophobic and has a very low surface energy of ~ 30 mJ/m2 (The surface energy for Si is ~
1400 mJ/m2; and the surface tension of water is 72 mJ/m2). These properties prevent
contaminants from accumulating on the surface and hence prevent the contamination of the tip
in the evaluation process. This method of using BOPP to check AFM tips AND to clean
contaminated tips was highlighted in April 1, 2001 issue of Analytical Chemistry.

3.2 Applications to Blind Tip Reconstruction


Considerable effort has been expended to mathematically extract the geometry of the tip
based solely on an algorithm derived from a given image. The method is known as blind
reconstruction. This methodology is based on an assumption that protrusions in the AFM image
represent the self-image of the tip, which is equivalent to the statement that sharper features on
the sample surface act as the probe to image the AFM tip. This method has proven useful and
successful in estimating tip geometry from an existing image, when appropriate samples were
chosen (i.e., some surface features on the sample are sharper than the tip). Once the tip
geometry is known, the tip effect may be subtracted from the original image through the
mathematical operation of erosion, also known as deconvolution. Dilation is another
mathematical operation, which adds a tip effect to an existing AFM image by “scanning” the
known tip across the “surface” of the image. This transformation appears useful in simulating
tip effect to a given image, because the mechanism of AFM can be regarded as a dilation
between the tip geometry and surface features of the sample.
We have found that a BOPP film is
suitable for checking tip performance
and for cleaning contaminated tips,
thus making it possible to collect
images of the same area of a BOPP
film surface before and after the tip
was cleaned. Therefore, the
difference between the two different
images is solely due to the
contamination of the tip. We took
advantage of our ability to collect
AFM images of the same area using
the same tip, in one instance,
contaminated and, in the other, after
being cleaned. Commercial software SPIP (Metrology Image ApS, Denmark) was used to
estimate the tip geometry using its “tip characterization module”, in which the blind
reconstruction algorithm is implemented. First we used blind reconstruction on the image
collected using the contaminated tip. Blind tip reconstruction allows one to extract the geometry
of the tip from a given image. Once we had estimated the geometry of the contaminated tip, we
used it to simulate the tip effect using the image collected using the cleaned tip. By comparing
the simulation result with the image collected using the contaminated tip we showed that the
blind reconstruction routine works well. Prior to this, there was no de facto method for testing
blind reconstruction algorithms.
Comparison of tip geometry from the blind reconstruction method and from scanning
electron microscopy (SEM) images has been made by Dongmoet al. [S. Dongmo, M. Troyon, P.
Vautrot, E. Delain, and N. Bonnet, J. Vac. Sci. & Technol. B 14, 1552 (1996)]. We provides a
simpler way to test blind reconstruction: comparison of AFM images collected in the same area
of the BOPP by clean and contaminated tips. If the estimation of the contaminated tip geometry
is reasonable, then one expects to be able to use the estimated tip geometry to dilate the image
collected using the clean tip to obtain an image resembling one collected using the contaminated
tip. Conversely, one can also determine if the deconvolution works by eroding the image
collected with the contaminated tip using the estimated tip geometry to see whether the result
resembles the image collected using the clean tip.
Because an AFM image
is a convolution of the
surface features and the tip
geometry, if neither of
them is known, there is no
way to know, on an
unknown sample, if the
image is dominated by the
surface features or the tip
effect. When the tip is
much sharper than the
surface features, it will
collect an image reflecting
the “true” surface features.
This is the reason why a
reference sample is
essential to check the tip
performance. It is
important to note that a tip
could be easily
contaminated or damaged depending on the chemical and mechanical properties of the sample
surface. Using electron microscopes one can evaluate the outlines of the tip shape from specific
directions, but it is difficult, if not impossible, to capture the three-dimensional geometry of the
tip. In combination with blind reconstruction, using BOPP film to check the tip performance
provides a simple and effective protocol to test the estimation of the tip geometry of a
contaminated tip. One can do this by comparing the image collected using the contaminated tip
with the image generated by dilating the image collected using the clean tip.
On the other hand, when the tip is much larger than the surface features, using such a tip to
scan the surface will result in an image that is merely a reflection of the geometry of the tip apex
itself. In this case, it is evident that the information about the surface features is physically lost.
Therefore, the erosion operation will not lead to the recovery of the “true” surface features,
though the mathematic operation may result in an image which is likely closer to the “true”
surface features. The degree of the recovery by the erosion operation is dependent on how
severely the tip is contaminated. One can imagine that different surface features can have
similar images if a large tip is used: they are dominated by the tip effect.
References
 H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Atomic Force Microscopy Study of Biaxially-
Oriented Polypropylene Films", J. Mater. Eng. Perform., 13, pp.451-460 (2004).
 H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Use of biaxially-oriented polypropylene film
for evaluating and cleaning contaminated atomic force microscopy probe tips: An application to
blind tip reconstruction", Rev. Sci. Instrum. 73, pp.3831-3836 (2002).
 H.-Y. Nie and N.S. McIntyre, "A simple and effective method of evaluating atomic force
microscopy tip performance", Langmuir 17, pp.432-436 (2001).

Return to Top
4. A general method to clean contaminated tips using
UV/ozone treatment
The tip can be contaminated during scanning some surfaces or just left in air as recognized
by the unstable and degraded images obtained by the tip. When the tip was in this condition, we
took out the tip for 5 minute treatment in UV/ozone. After that, the imaging condition became
stable and images were improved largely. Therefore, the UV/ozone treatment is effective to
clean the tip and hence opened a way of recycle-using probe tips.

The wavelength of UV light from a mercury lamp is mainly 253.7 nm; with a much lower
percentage at 184.9 nm. Photons with those two wavelength are effective for cleaning organic
contaminants.

Although ozone can be generated by irradiating oxygen (air) with short wavelength light
(184.9 nm; photon energy at this wavelength is 6.70 eV or 154.59 kcal/mol), a separate ozone
source (such as an ozone generator) is required to provide enough ozone concentration to clean
the contaminants fast enough. What is really doing the cleaning job in short time is the atomic
oxygen, which is produced by the decomposition of ozone in the presence of UV light (253.7
nm; photon energy at this wavelength is 4.89 eV or 112.66 kcal/mol). This atomic oxygen
oxidizes organic contaminants to form volatile molecules. Meanwhile the UV light also has an
effect to excite the contaminant molecules to make them more reactive with ozone and/or
atomic oxygen. Ozone itself is reactive with organic contaminant, therefore, ozone alone is also
able to clean organic contamination but will take much longer time than UV/ozone combined.
Return to Top
5. Concluding remarks
SPM techniques have been developed extremely fast. It is a promising tool and a base for the
new nano science and technology. More and more researchers in many different fields are using
SPM. Some want to develop a technology based on SPM to fabricate nano-scale devices. Others
are discovering knowledge in physics, chemistry, biology and materials science on nano- and/or
meso-scale. SPM promises to provide us new and exciting discovers in surface science and
technology, physics, chemistry, materials science and biological technology. This direction is
clear if one notes that news media and government agencies are actively involved in reporting
and supporting discoveries and development in nanotechnology. The base for such a new field
is, of course, SPM.

Return to Top
References
G. Binnig, H. Rohrer, Ch. Gerber and E. Weibel, Appl. Phys. Lett. 40, 178 (1982).
G. Binnig, H. Rohrer, Ch. Gerber and E. Weibel, Phys. Rev. Lett. 49, 57 (1982).
G. Binnig, C.F. Quate and Ch. Gerber, Phys. Rev. Lett. 56, 930 (1986).
R. Wiesendanger, Scanning Probe Microscopy and Spectroscopy, Cambridge University Press,
1994.
J.N. Israelachvili, Intermolecular and Surface Forces, 2nd ed., Academic Press, 1992.
AFM Manufacturers
Park Systems
Veeco Instruments | Seiko Instruments | Agilent Tchnologies
Asylum Research | NT-MDT | Nanotec Electronica
JPK Instruments | Surface Imaging Systems | Anfatec
Concentris (cantilever sensor technology)

Publication list on SPM work


33) H.-Y. Nie, J.T. Francis, A.R. Taylor, M.J. Walzak, W.H. Chang, D.F. MacFabe and W.M. Lau, "Imaging
subcellular features of a sectioned rat brain using time-of-flight secondary ion mass spectrometry and scanning
probe microscopy", Appl. Surf. Sci. 255, 1079-1083 (2008).
32) H.-Y. Nie, N.S. McIntyre, W.M. Lau and J.M. Feng, "Optical properties of octadecylphosphonic acid self-
assembled monolayer on a silicon wafer", Thin Solid Films 517, 814-818 (2008).
31) H.-Y. Nie, N.S. McIntyre and W.M. Lau, "Nanolithography of a full-coverage octadecylphosphonic acid
monolayer spin coated on a Si substrate", Appl. Phys. Lett. 90, 203114 (2007).
30) H.-Y. Nie, N.S. McIntyre and W.M. Lau, "Selective removal of octadecylphosphonic acid (OPA) molecules
from their self-assembled monolayers (SAMs) formed on a Si substrate", Journal of Physcis: Conference Series 61,
869-873 (2007).
29) H.-Y. Nie and N.S. McIntyre, "Unstable amplitude and noisy image induced by tip contamination in dynamic
force mode atomic force microscopy", Rev. Sci. Instrum. 78, 023701 (2007).
28) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Growth and properties of complete monolayer films of
octadecylphosphonic acid (OPA) on oxidized aluminum surfaces", ATB Metallurgie 45, 564-568 (2006).
27) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Scratch resistance anisotropy in biaxially oriented polypropylene
and poly(ethylene terephthalate) films", Appl. Surf. Sci., 253, pp.2320-2326 (2006).
26) H.-Y. Nie, M.J. Walzak and N.S. McIntyre,"Delivering Octadecylphosphonic Acid Self-Assembled
Monolayers on a Si Wafer and Other Oxide Surfaces", J. Phys. Chem. B, 110, pp.21101-21108 (2006).
25) J.T. Francis, H.-Y. Nie, N.S. McIntyre and D. Briggs, "ToF-SIMS Investigation of Octadecylphosphonic Acid
Monolayers on a Mica Substrate", Langmuir, 22, pp.9244-9250 (2006).
24) N.S. McIntyre, H.-Y. Nie, A.P. Grosvenor, R.D. Davidson and D. Briggs, "XPS studies of
octadecylphosphonic acid (OPA) monolayer interactions with some metal and mineral surfaces", Surf. Interf. Anal.
37, pp.749-754 (2005).
23) H.-Y. Nie, D.J. Miller, J.T. Francis, M.J. Walzak and N.S. McIntyre, "Robust self-assembled
octadecylphosphonic acid monolayers on mica substrate", Langmuir 21, pp.2773-2778 (2005).
22) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Atomic Force Microscopy Study of Biaxially-Oriented
Polypropylene Films", J. Mater. Eng. Perform., 13, pp.451-460 (2004).
21) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Use of biaxially-oriented polypropylene film for evaluating and
cleaning contaminated atomic force microscopy probe tips: an application to blind tip reconstruction", Rev. Sci.
Instrum. 73, pp.3831-3836 (2002).
20) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Bilayer and odd-numbered multilayers of octadecylphosphonic
acid formed on a Si substrate studied by atomic force microscopy", Langmuir 18, pp.2955-2958 (2002).
19) H.-Y. Nie and N.S. McIntyre, "A simple and effective method of evaluating atomic force microscopy tip
performance", Langmuir 17, pp.432-436 (2001).
18) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Atomic force microscopy study of UV/ozone treated
polypropylene films", Polymer Surface Modification: Relevance to Adhesion, Vol.2, K.L. Mittal Ed., VSP (Utrecht,
The Netherland), pp.377-392 (2000).
17) H.-Y. Nie, M.J. Walzak and N.S. McIntyre, "Draw-ratio-dependent morphology of biaxially-oriented
polypropylene films as determined by atomic force microscopy"(Vol. and page in Ref. 3 and 14 should be
exchanged), Polymer 41, pp.2213-2218 (2000).
16) H.-Y. Nie, M.J. Walzak, B. Berno and N.S. McIntyre, "Microscopic stripe formation and adhesion force
increase introduced by local shear-stress deformation of polypropylene film", Langmuir 15, pp.6484-6489 (1999).
15) H.-Y. Nie, M.J. Walzak, N.S. McIntyre and A.M. EL-Sherik, "Applications of lateral force imaging to enhance
topographic features of polypropylene film and photo-cured polymers", Appl. Surf. Sci. 144-145, pp.633-637
(1999).
14) H.-Y. Nie, M.J. Walzak, B. Berno and N.S. McIntyre, "Atomic force microscopy study of polypropylene
surfaces treated by UV and ozone: modification of morphology and adhesion force", Appl. Surf. S ci. 144-145,
pp.627-632 (1999).
13) H.-Y. Nie and J. Masai, "Surface potentials on Pd/GaAs contacts studied using scanning probe microscopy",
Appl. Phys. A 66, pp.s1059-s1062 (1998).
12) H.-Y. Nie, K. Horiuchi, H. Yamauchi and J. Masai, "Local surface potential measurement of Pd/GaAs contact
and anodized aluminum films using scanning probe microscopy", Nanotechnology 8, pp.A24-A31 (1997).
11) H.-Y. Nie, M. Motomatsu, W. Mizutani and H. Tokumoto, "Observation of modification and recovery of local
properties of polyethylene oxide", J. Vac. Sci. Technol. B15, pp.1388-1393 (1997).
10) M. Motomatsu, T. Takahashi, H.-Y. Nie, W. Mizutani and H. Tokumoto, "Microstructure study of acrylic
polymer-silica nanocomposite surface by scanning force microscopy", Polymer 38, pp.177-182 (1997).
9) M. Motomatsu, W. Mizutani, H.-Y. Nie and H. Tokumoto, "Surface structure of a fluorinated thiol on Au(111)
by scanning force microscopy", Thin Solid Films 281-282, pp.548-551 (1996).
8) M. Motomatsu, H.-Y. Nie, W. Mizutani and H. Tokumoto, "Surface morphology study of poly (ethylene oxide)
crystals by scanning force microscopy", Polymer 37, pp.183-185 (1996).
7) M. Motomatsu, H.-Y. Nie, W. Mizutani and H. Tokumoto, "Scanning force microscopy application to polymer
surfaces for novel nano-scale surface characterization", Thin Solid Films 273, pp.304-307 (1996).
6) H.-Y. Nie, M. Motomatsu, W. Mizutani and H. Tokumoto, "Local elasticity measurement on polymers using
atomic force microscopy", Thin Solid Films 273, pp.143-148 (1996).
5) H.-Y. Nie, M. Motomatsu, W. Mizutani and H. Tokumoto, "Local modification of elastic properties of
polystyrene-polyethyleneoxide blend surfaces", J. Vac. Sci. Technol. B13, pp.1163-1166 (1995).
4) M. Motomatsu, W. Mizutani, H.-Y. Nie and H. Tokumoto, "Lateral force measurements on phase separated
polymer surfaces", in Proceedings of Forces in Scanning Probe Microscopies, edited by H.-J. Guntherodt et al.,
ASI E286 (Kluwer Academic, Dordrecht, 1995), 331-336.
3) M. Motomatsu, H.-Y. Nie, W. Mizutani and H. Tokumoto, "Local properties of phase-separated polymer
surfaces by force microscopy", Jpn. J. Appl. Phys. 33, Part 1, pp.3775-3778 (1994).
2) H.Y. Nie, W. Mizutani and H. Tokumoto, "Au(111) reconstruction observed by atomic force microscopy with
lateral force detection", Surf. Sci. 311, pp.L649-L654 (1994).
1) H.Y. Nie, T. Shimizu and H. Tokumoto, "Atomic force microscopy study of Pd clusters on graphite and mica",
J. Vac. Sci. Technol. B12, pp.1843-1846 (1994).

spacing
Examples of AFM images
01
AFM related infomation
Analysis Services
00000
Polymer AFM
Introduction on AFM (PDF)
Publication
Last modified on May 16, 2010 Talks and posters
H.-Y. Nie
Surface Science Western, Room G-1 WSC
| Return to Top |
The University of Western Ontario spacing
London, Ontario N6A 5B7, Canada Since April, 2000
Phone: (519) 661-2173; Fax: (519) 661-3709
Since March 2002

Вам также может понравиться