You are on page 1of 17

Atomic Force Microscopy (AFM) UNIT 2C.

2
1 2
Andreea Trache and Gerald A. Meininger
1
Department of Systems Biology & Translational Medicine, College of Medicine, Texas
A&M Health Science Center, College Station, Texas
2
Dalton Cardiovascular Research Center, University of Missouri, Columbia, Missouri

ABSTRACT
The atomic force microscope (AFM) is an important tool for studying biological sam-
ples due to its ability to image surfaces under liquids. The AFM operates by physical
interaction of a cantilever tip with the molecules on the cell surface. Adhesion forces
between the tip and cell surface molecules are detected as cantilever deflections. Thus,
the cantilever tip can be used to image live cells with atomic resolution and to probe
single molecular events in living cells under physiological conditions. Currently, this is
the only technique available that directly provides structural, mechanical, and functional
information at high resolution. This unit presents the basic AFM components, modes of
operation, useful tips for sample preparation, and a short review of AFM applications
in microbiology. Curr. Protoc. Microbiol. 8:2C.2.1-2C.2.17.  C 2008 by John Wiley &

Sons, Inc.
Keywords: AFM r imaging r force curves

INTRODUCTION
Understanding the function of microbial cell surfaces requires knowledge of their struc-
tural and physical properties. Most of the classical methods used to investigate these
structures (e.g., electron microscopy, X rays) require cell drying, which may compro-
mise the accuracy of the analysis by denaturing the surfaces (Dufrêne, 2002, 2003).
Because of microorganisms’ small size, the information is generally obtained from a
large number of cells and not at the individual cell level. Thus, there is a need for new,
nondestructive technologies capable of probing single cell surfaces at high resolution.
The atomic force microscope (AFM) has emerged as a nondestructive technique, offer-
ing resolution comparable to electron microscopy but allowing the study of biological
material in its native environment. AFM was invented in 1986 by Binnig et al. (1986) as
a novel technique that allows imaging of surfaces and force measurements at the atomic
scale, whether the sample is in air or under fluid. This last property makes it very attrac-
tive as a tool for biological applications for live cell imaging (Radmacher et al., 1992;
Lal and John, 1994) and single molecule force spectroscopy (Willemsen et al., 2000;
Wojcikiewicz et al., 2004; Trache et al., 2005), providing structural, mechanical, and
functional information with high resolution under physiological conditions. Biologists
are using the AFM in an increasing number of creative ways, and the referenced literature
is vast (see Key References). This unit presents the basic AFM components and modes of
operation, gives useful tips for sample preparation, and provides a short review of AFM
applications in microbiology.

PRINCIPLE OF OPERATION
The principle of operation of the AFM is very similar to that of a stylus profilometer: a
sharp cantilever tip interacts with the sample surface, sensing the local forces between
the molecules of the tip and sample surface (Fig. 2C.2.1). This instrument is not a
conventional microscope that collects and focuses light. The word microscope has been

Microscopy

Current Protocols in Microbiology 2C.2.1-2C.2.17, February 2008 2C.2.1


Published online February 2008 in Wiley Interscience (www.interscience.wiley.com).
DOI: 10.1002/9780471729259.mc02c02s8 Supplement 8
Copyright C 2008 John Wiley & Sons, Inc.
Figure 2C.2.1 Schematic illustration of the AFM system. A flexible cantilever with a tip at the
end is rigidly connected to an xyz piezoelectric element. The optical lever consists of a laser diode
beam that is focused on the back of the cantilever and bounces off, reaching the quadrant photode-
tector. In contact mode the xyz position of the cantilever is given by recording the corresponding
position of the piezoelectric element, and the detector reading is proportional to the cantilever
deflection. In tapping mode the AFM cantilever is vibrated at its resonant frequency under an
external electrical excitation (dashed line).

associated with this instrument because it is able to measure microscopic features of the
sample. The most characteristic property of the AFM is that the images are acquired
by “feeling” the sample surface without using light. A common analogy is to imagine
a blindfolded person who perceives the characteristics of a surface just by touching it.
In this way, not only the sample topography can be recorded with atomic resolution,
but also material characteristics and the eventual strength of the interaction between the
sample surface and the AFM tip. Due to the fact that no light is involved in acquiring the
sample properties, the resolution achieved by the AFM is not limited by the wavelength
of the radiation that investigates the sample, as in classical light microscopy. Thus, the
AFM reaches a resolution far below the diffraction limit offered by light microscopy,
being limited only by the tip radius and the spring constant of the cantilever.

The main components of an AFM (Gad and Ikai, 2001) consist of the AFM probe (a
sharp tip mounted on a soft cantilever), the optical lever that allows for measuring the
cantilever deflections, the feedback loop that allows for monitoring the interaction forces
between the molecules on the tip with the ones on the cell surface, the piezoelectric
scanner that moves the tip in respect with the sample in a 3D pattern, and a conversion
system from raw data acquired by the instrument into an image or other useful display.
A short overview of the system components is presented below.

The most sensitive part of an AFM is the tip that interacts directly with the sample
surface. The most common probes are constructed from silicon or silicon nitride using
microfabrication techniques (Binnig et al., 1987). Each probe can have multiple integrated
cantilevers (Fig. 2C.2.2). The properties and dimensions of the cantilever and tip play a
major role in determining the sensitivity and resolution of the AFM (Morris et al., 1999).
For special applications the pyramidal tip can be replaced with glass or polystyrene beads
with a diameter of 1 to 5 µm. The AFM tip must be chosen carefully depending on the
specific application. In general, the cantilever should be soft enough to be deflected at
Atomic Force
Microscopy

2C.2.2
Supplement 8 Current Protocols in Microbiology
Figure 2C.2.2 Scanning electron microscope images of AFM cantilevers. (A) Typical AFM probe
for force measurements (magnification 45×). At the very end of the cantilever is a pyramid-shaped
tip with a 4-µm base, 35◦ half angle of the tip, and a radius of 50 nm (inset, magnification 16,000×).
(B) Typical AFM probe for mechanical stimulation measurements (magnification 10,000×). The
cantilever is V-shaped, with a spring constant of ∼0.06 N/m and a 2-µm glass bead attached at
its end (inset, magnification 10,000×). Images were taken at the Microscopy & Imaging Center of
TAMU, College Station, Texas. Reproduced from Trache and Meininger (2005) with permission of
the International Society for Optical Engineering.

Table 2C.2.1 Most Common Cantilever Parameters

Cantilever material Silicon nitridea Silicona

Spring constant (k) 0.58, 0.32, 0.12, 0.06 20-100 N/m


Resonant frequency 200-400 kHz
Cantilever V-shaped Single beam
configuration
Reflective coating gold Uncoated, aluminum
b
Tip shape Pyramid Sphere Pyramid
glass/polystyrene
Tip radius of 20-60 nm 0.6-45 µmb 5-10 nm
curvature/diameter
Sidewall angles 35◦ on all four sides NA 35◦ on all four sides
a Adapted from SPM Training Notebook (2003) with permission of Veeco Instruments.
b Parameters from Novascan Technologies.
Abbreviation: NA, not applicable.

very small forces, and the tip radius should be comparable with the features of interest.
Table 2C.2.1 presents the properties of the most common cantilevers.

Use of the optical lever is the most common technique for detecting the cantilever
deflection. A laser beam coming from a laser diode is reflected off the back of the
cantilever onto a photodetector. The photodetector consists of a split photodiode that
senses the change in light intensity from the reflected beam due to changes in the
deflection of the cantilever following tip-sample interaction. The sensitivity of such a
system is in the range of 0.1 nm (Meyer and Amer, 1990). Microscopy

2C.2.3
Current Protocols in Microbiology Supplement 8
The AFM scanners are made from piezoelectric material, which expands or contracts
proportionally with an applied voltage. The piezoelectric scanner can move very precisely
with very good reproducibility for small displacements, but for displacements >70% of
the full-scale displacement, the piezoelectric response is nonlinear. The commercially
available AFM featuring closed-loop feedback systems independently monitors the scan-
ner movement and corrects its motion for nonlinearity. There are two types of scanner
configurations: scanned tip AFM (where the piezoelectric scanner is rigidly attached
to the probe and is moved over the sample surface, which stands still) and scanned
sample AFM (where the scanner is attached to the sample and is moved under the tip).
Both designs could be integrated with optical microscopes or used as stand-alone AFM
systems.

A dual monitor computer interface allows for display of data (images or force curves) on
the display monitor and the parameters of the system on the control monitor.

OPERATION MODES OF AFM


Contact Mode Imaging
The most common imaging mode is the constant force mode. In this case, the AFM tip
is brought in contact with the sample surface and set to scan the sample in an xy raster
pattern. The feedback loop maintains a constant deflection (force) of the cantilever with
respect to the sample surface by moving the z scanner for each xy coordinate. This change
in z axis corresponds to the topographical height of the sample at each given point. The
height image preserves the true height information of the sample. Also, a “deflection”
image can be recorded by monitoring the cantilever deflection from a straight line. This
image is not recorded at constant force, the tall features in the image representing regions
of higher force (i.e., higher degree of cantilever bending). This type of imaging loses the
true height information, but it presents more fine details of the sample than the height
image because the feedback loop response is faster for correcting the position of the
small cantilever, in comparison with correcting the position of the piezoelectric element.

Tapping Mode Imaging


For very soft samples, contact mode might not be the best choice due to the friction
between the AFM tip and the sample surface. In tapping mode, the cantilever vibrates
at its resonant frequency (bounces up and down) under an external electrical excitation.
While rastering the sample in xy the AFM tip briefly touches the sample at the bottom
of each swing, producing a decrease in the oscillation amplitude. Similar to the constant
force in contact mode, the feedback loop keeps this decrease in oscillation amplitude at
a constant value by moving the piezoelectric tube, and a height image can be recorded
(Putman et al., 1994). In addition, a “phase” image can be recorded in tapping mode.
When the tip touches the sample at the bottom of its swing, the phase of oscillation is
disturbed, inducing a phase difference between the tip and the electrical oscillator that is
driving it (Morris et al., 1999). The contrast of a phase image is directly dependent on the
elastic properties of the sample. This is not a preferred mode for imaging cells because
the elastic properties of the cell surface do not vary enough from point-to-point to offer
a good phase-contrast image.

Dynamic Force Spectroscopy Mode


For force measurements the AFM is operated in force mode. The piezoelectric scanner
is set to drive the cantilever to touch and retract over a predefined distance in the z axis at
Atomic Force a fixed xy position, in such a way that the z axis movement of the piezoelectric element
Microscopy and the deflection signal from the cantilever are recorded in a force curve. In this way,
2C.2.4
Supplement 8 Current Protocols in Microbiology
Figure 2C.2.3 Diagram of generic force curves. Approach (black thin line) and retraction force
curves with (black thick line) and without (middle line) adhesions are presented. The x axis
represents the piezoelectric element displacement, and the y axis represents the force (see text
for details).

measurements of the relative stiffness of the cell surface (approach curve) and adhesion
force measurements between the biologically functionalized AFM tip and the cell surface
(retraction curve) are acquired (Fig. 2C.2.3). When the probe is extended towards the
cell surface (a), a contact point is established with the cell (b), and thereafter the cell
surface is indented. Due to cell stiffness, further probe extension causes an opposing
force of increasing magnitude to be generated along with increasing indentation in the
cell membrane (b to c). The upward deflection of the cantilever as it bends in response
to this force, results in an increasing deflection signal. When the probe retracts from
the sample (c to d), the force between probe and sample gradually decreases until the
cantilever returns to the original position, at which time the deflection signal returns to
the original value (a). However, if adhesions occurred between the probe and sample
surface, the force causes the cantilever to bend downward, and the deflection signal will
be lower than the original value (e to g). When all adhesions are broken, the cantilever
returns to the original position (a), and the deflections (unbinding of adhesion events) are
recorded on the retraction force curve. The adhesion force is calculated as the product of
the deflection height associated with the unbinding event and the spring constant of the
cantilever.

On soft samples such as cells, the force application results in an indentation of the
cell that is a measure of the local elastic properties of the cell membrane. Due to this
elasticity of the cell membrane, only a portion of the piezoelectric element movement
causes deflection of the probe; the remainder results in indentation of the cell membrane.
The membrane indentation part of the approach curve could be analyzed using different
models (Sneddon, 1965; Costa and Yin, 1999) in order to determine the local values of
the Young modulus of elasticity as a measure of the local apparent elasticity of the cell
membrane.

PRACTICAL GUIDELINES
Although AFM has great potential in microbiological studies, it is important to understand
that accurate data collection is sometimes difficult due to the technical limitations associ-
ated with the technique itself. Images produced by AFM always represent a convolution
of the tip shape with the sample features. It is important to correlate the tip geometrical Microscopy

2C.2.5
Current Protocols in Microbiology Supplement 8
Figure 2C.2.4 Schematics of tip artifacts. (A) The smaller the tip radius of curvature, the smaller
the sample feature that can be resolved. A dull or dirty (B) tip will affect the image profile. Adapted
from SPM Training Notebook (2003) with permission of Veeco Instruments.

features with the expected features of the sample to obtain the best images. A tip that is
too large will not be able to record features in the sample smaller than the tip parameters
(Fig. 2.C2.4A). Also, a contaminated, broken, or blunt tip will introduce distortions into
an image, affecting its sharpness or even introducing noise, with the geometry of the tip
dominating the geometry of the sample (Figure 2.C2.4B). A detailed description of AFM
artifacts is presented at http://www.pacificnanotech.com/afm-artifacts single.html.

Sample Immobilization
One of the most important aspects of imaging live cells under physiological conditions
using the AFM is sample immobilization. If the sample is not firmly attached to the
bottom of cell culture dish, the AFM tip can easy lift the sample. In this case the tip
will break or get contaminated in the process, and no measurements can be performed.
Most live cells (e.g., smooth muscle cells, endothelial cells, fibroblasts) attach well
to classical substrates like polystyrene or glass. Some cells (e.g., myocytes) can be
problematic for AFM measurements due to their intrinsic movement or poor attachment
to the substrate. Coating the cell culture dish with gelatin or an extracellular matrix
protein (e.g., fibronectin, laminin) can considerably enhance the cell attachment. Live cell
imaging can be performed at room temperature or 37◦ C depending on the experimental
requirements. For imaging single bacteria, yeast, or fungal cells under physiological
conditions, a concentrated cell suspension is absorbed through an isopore polycarbonate
membrane with a pore size comparable to the cell size (Kasas and Ikai, 1995; Dufrêne
et al., 1999). Imaging of reconstituted microbial layers or DNA is performed by absorbing
the layers of interest on a flat substrate, i.e., fresh cleaved mica or glass (Müller et al.,
1997; Colton et al., 1998). Other immobilization methods involve air-drying the sample
Atomic Force (a droplet of a concentrated cell suspension is placed on a mica or glass substrate and
Microscopy

2C.2.6
Supplement 8 Current Protocols in Microbiology
allowed to dry in air; Amro et al., 2000) or covalently bonding bacteria to silanized glass
slides (Camesano et al., 2000).

Labeling AFM Tips


For force spectroscopy measurements, the AFM tips are typically biofunctionalized with
ligands of interest to allow the study of ligand-receptor or protein-protein interaction.
This process consists of cross-linking a protein of interest on the AFM tip that has the
role of ligand, while its specific receptor is situated on the cell surface. The conjugation
can be performed in-laboratory or by using specialized services for more complicated
conjugation processes. Two simple ways of cross-linking fibronectin on the AFM tip are
presented here. These methods can be applied for other proteins as well.

Coating silicon nitride cantilever probes with fibronectin


For adhesion force measurements, the silicon nitride cantilevers probes (ThermoMicro-
scope) can be coated (Lehenkari and Horton, 1999) with 1 mg/ml fibronectin (FN).
Polyethylene glycol (Sigma) at 10 mg/ml is used to cross-link fibronectin onto probes
at room temperature (Hinterdorfer et al., 2000). After the tip is mounted on the glass
holder and washed, it is incubated 5 min with polyethylene glycol, washed five times
with deionized water, and then incubated 1 min with fibronectin. The tip is then washed
again five times with phosphate-buffered saline (APPENDIX 2A) and mounted on the AFM
scanner.

Coating modified silicon nitride cantilevers with avidin-biotin/fibronectin


A more robust coating that can be used for mechanical cell stimulation consists of
avidin (Sigma) cross-linked with biotin/fibronectin (Pierce) onto functionalized glass
beads attached to silicon nitride cantilevers (Novascan Technologies). The cantilevers are
customized by Novascan Technologies, such that glass beads are glued to the cantilever
and then prefunctionalized with biotin. The probes are first incubated 5 min with avidin
(1 mg/ml), washed with phosphate-buffered saline five times, and then incubated 5 min
with biotin/fibronectin. The tip is washed again with phosphate-buffered saline five times.

The coating should be performed only at the very end of the cantilever to avoid altering
its spring constant that is assumed to be unchanged after protein labeling.

AFM APPLICATIONS SPECIFIC TO MICROBIOLOGY


AFM is the only technique available that directly provides structural, mechanical, and
functional information at high resolution of cells under physiological conditions. In
comparison to electron microscopy, which requires cell fixation and drying, AFM of-
fers the possibility of acquiring high-resolution data in living specimens in real time.
Tables 2C.2.2 and 2C.2.3 provide examples of AFM studies relevant to microbiology,
focusing on imaging cell surface layers using different immobilization procedures and
imaging living cells under physiological conditions (Dufrêne, 2002), respectively. To
avoid air-drying techniques that may induce protein or cellular surface denaturing, the
physiological imaging techniques are preferred when appropriate. Table 2C.2.4 shows a
list of microorganisms for which molecular interactions or local elasticity measurements
have been performed. Combining force measurements with high-resolution imaging of
cells will be an important challenge, as this will make possible to correlate molecular
interactions with structural changes. The examples presented here are just a few of many
applications that can be developed in applying AFM to microbiology.

Microscopy

2C.2.7
Current Protocols in Microbiology Supplement 8
Table 2C.2.2 Imaging the Ultrastructure of Cell Surface Layersa

Organism Sample Immobilization procedure Observationsb Reference

Bacillus coagulans S-layer Covalent linkage to Oblique lattice; r ≈ 10 nm Ohnesorge et al.


glass/mica (study of antibody binding) (1992)
Recrystallization on Oblique lattice; r = 1-2 nm Pum and Sleytr (1996)
silanized silicon (agreement with electron
microscopy)
Bacillus sphaericus S-layer Covalent linkage to Square lattice; r ≈ 12 nm Ohnesorge et al.
glass/mica (agreement with electron (1992)
microscopy)
Recrystallization on Square lattice; r = 1-2 nm Pum and Sleytr (1996)
silanized silicon (agreement with electron
microscopy)
Recrystallization on Square lattice; r = 1-2 nm Wetzer et al. (1997)
supported lipid bilayers (novel S-layer/lipid bilayer
structure)
Bacillus S-layer Recrystallization on Oblique lattice; r ≈ 1.5 nm Pum and Sleytr (1995)
stearothermophilus various silicon surfaces (study of recrystallization
process)
Deinococcus Hexagonally Covalent linkage to glass Ring-shaped hexamers; r = 1 Karrasch et al. (1994)
radiodurans packed nm (correlation with electron
intermediate microscopy)
layer
Adsorption on mica Conformational change of Müller et al. (1996)
central pores
Adsorption on HOPG Tapping mode; r = 1-1.5 nm Möller et al. (1999)
Halobacterium Purple Adsorption on mica, Hexagonal symmetry; r = Butt et al. (1990)
laobium membrane silanized glass, and 1.1 nm
supported lipid bilayers
Halobacterium Purple Adsorption on mica Conformational changes; r < 1 Müller et al. (1997b )
salinarum membrane nm
Tapping mode; r = 1-1.5 nm Möller et al. (1999)
Trigonal lattice; r = 0.5 nm Müller et al. (2000)
(antibody labeling;
complementarity with X-ray
and electron crystallography)
Escherichia coli Porin OmpF Assembly on mica in the Porin trimers; r = 1 nm Schabert et al. (1995)
presence of lipids (comparison with X-ray
structure; detection of two
conformations)
Voltage and pH-dependent Müller and Engel
conformational changes (1999)
Escherichia coli Aquaporin Z Assembly on mica in the Tetramers; p42(1)2 and p4 Scheuring et al.
presence of lipids symmetry, r<1 nm (proteolytic (1999)
cleavage force-induced
conformational changes)
a Reproduced from Dufrêne (2002) with permission of the American Society for Microbiology.
b r-values are lateral resolutions determined directly from the images or after image processing.

Atomic Force
Microscopy

2C.2.8
Supplement 8 Current Protocols in Microbiology
Table 2C.2.3 Imaging the Surface Structure of Living Cellsa

Organism Observations Reference

Saccharomyces cerevisiae Low resolution, bud scars Kasas and Ikai (1995);
also see Figure 2C.2.8
In situ investigation of cell growth Gad and Ikai (1995)
process
High resolution; Dufrêne (2002)
smooth surface
Phanerochaete chrysosporium High resolution; 10 nm; Dufrêne et al. (1999)
rodlet structures, changes upon
germination
Aspergillus oryzae High resolution; 10 nm; Van der Aa et al. (2001)
rodlet structures, changes upon
germination
Lactococcus lactis Sample deformation and imaging Boonaert et al. (2002)
artifacts (holes, grooves)
Streptococcus salivarius Smooth surface vs surface Van der Mei et al. (2000)
appendages (fibrils)
Pseudomonas aeruginoas Surface morphology of biofilms; Steele et al. (1994)
extracellular polymeric substances
Pseudomonas putida Surface morphology of biofilms; Auerbach et al. (2000)
extracellular polymeric substances
a Reproduced from Dufrêne (2002) with permission of the American Society for Microbiology.

Investigating Actinomyces Species


Tang et al. (2004) used AFM to image and directly detect interactive forces between
the AFM tip and Actinomyces species that are predominant early colonizers of the oral
cavity and prime mediators of interbacterial adhesion. Figure 2C.2.5 shows the two-
dimensional and three-dimensional reconstruction of a height image of A. naeslundii
genospecies 1 adherent to mica substrate. In their paper, several sets of experiments were
performed in which the interaction forces between the AFM tip and different locations
on the same or different Actinomyces species were measured. This study was the first to
quantitatively evaluate interactive forces at cell surfaces of Actinomyces species at the
nanoNewton (nN) force range.

Imaging Tipula Iridescent Virus


The largest virus investigated by AFM is the insect virus tipula iridescent virus (TIV).
Figure 2C.2.6 shows such a virus, dried from water onto a mica substrate, forming
hexagonal arrays that have some degree of crystalline order. Although not properly
crystalline, the constraints introduced by association into an array make it possible
to measure center-to-center distances that agree well with height measurements and
dimensions determined by other methods (Williams, 1998; Kuznetsov et al., 2001).

Imaging Paramecium Bursaria Chlorella Virus Type 1 (PBCV-1)


Paramecium bursaria chlorella virus type 1 (PBCV-1) is a member of a large group
of viruses that infect certain unicellular, eukaryotic, chlorella-like green algae and is
common in fresh water bodies worldwide (Van Etten, 2000). Kuznetsov et al. (2005)
performed an AFM study to complement the electron microscopy data for this type of
virus. AFM can reveal structural anomalies and individual defects of single particles. Microscopy

2C.2.9
Current Protocols in Microbiology Supplement 8
Table 2C.2.4 Probing Cell Surface Properties, Molecular Interactions, and Elasticitya

Organism Method Results References

Escherichia coli Force-distance curves Attractive hydrophobic interactions; Lower et al. (2000),
with cell-coated probes repulsive steric interactions; Ong et al. (1999),
hydrophobic and electrostatic Razatos et al. (1998)
interactions
Shewanella oneidensis Force-distance curves Adhesion between cells and mineral Lower et al. (2001)
with cell-coated probes surfaces; adhesion peaks suggested
to reflect unfolding of a an outer
membrane protein
Burkholderia cepacia, Force-distance curves Electrostatic and steric interactions Camesano and Logan
Pseudomonas putida with silicon nitride probes (2000)
Phanerochaete Force-distance curves Hydrophobic and hydration Dufrêne (2000)
chrysosporium with chemically interactions, mapping of surface
functionalized probes hydrophobicity
Magnetospirillum Approach force curves Elasticity of cell wall, estimate of Arnoldi et al. (1998)
gryphiswaldense turgor pressure Arnoldi et al. (2000)
Streptococcus salivarius Approach force curves Softness of surface fibrils Van der Mei et al.
(2000)
Methanospirillum hungatei Depression technique Elasticity of proteinacious sheath Xu et al. (1996)
Escherichia coli, Depression technique Elasticity of murein sacculi Boulbitch et al. (2000)
Pseudomonas aeruginosa Yao et al. (1999)
Deinococcus radiodurans Retraction force curves Stretching single hexagonally packed Müller et al. (1999)
intermediate protomers, images of
single-molecule defects
Aspergillus oryzae Retraction force curves Stretching surface macromolecules Van der Aa et al. (2001)
a Reproduced from Dufrêne (2002) with permission of the American Society for Microbiology.

Figure 2C.2.7 shows high-magnification images of individual virions that exhibit a sur-
face appearing to be an open network with large spaces of almost hexagonal shape.

Visualizing Dynamic Events of Saccharomyces cerevisiae


Figure 2C.2.8 represents a three-dimensional reconstruction of an AFM height image
under fluid from a Saccharomyces cerevisiae cell trapped in a porous membrane (Dufrêne,
2002). The cell surface is smooth and shows a circular protrusion attributed to a bud scar.
The main advantage of using AFM at low resolution (as done here) is that dynamic events
such as the cell growth and budding process (Gad and Ikai, 1995; Kasas and Ikai, 1995)
and the change in cell surface morphology resulting from treatment with external agents
(e.g., enzymes or antibiotics) can be explored in real time.

Force Spectroscopy Measurements on S. cerevisiae at the NanoNewton Level


As mentioned earlier, AFM can also be used in the force spectroscopy mode to measure
molecular interactions and physical properties by functionalizing the AFM tip with
chemical coatings (Ahimou et al., 2002; Dufrêne, 2003). Such an example is shown
in Figure 2C.2.9 where the AFM tip functionalized with carboxyl groups maps the
electrostatic properties of S. cerevisiae at the nanometer level. Changes in adhesion
forces were measured as a function of pH, providing information on local isoelectric
points.
Atomic Force
Microscopy

2C.2.10
Supplement 8 Current Protocols in Microbiology
Figure 2C.2.5 Height and deflection images of A. naeslundii genospecies 1 (ATCC #12104) adherent to the
mica substrate after a 5-day period of anaerobic growth. The shape of the outer bacterial surface and the typical
diptheroidal arrangements (V, Y, or T forms and palisades) of Actinomyces species are seen in these images.
Reprinted from Archives of Oral Biology, vol. 49. Tang, G., Yip, H.K., Samaranayake, L.P., Chan, K.Y., Luo, G., and
Fang, H.H.P. Direct detection of cell surface interactive forces of sessile, fimbriated, and nonfimbriated Actinomyces
spp. using atomic force microscopy. pp. 727-738. Copyright 2004, with permission from Elsevier.

Figure 2C.2.6 When virus particles are clustered into two-dimensional arrays on the mica substrate, they are more firmly
immobilized and, therefore, yield better AFM images. The progressively magnified hexagonal arrangement seen here is
of the insect tipula iridescent virus (TIV). The little capsid structure is visible, and the polygonal structure of the capsid is
evident. Reproduced from Kuznetsov et al. (2001) with permission of the Society for General Microbiology.

Effect of BMS Biosurfactant on Dental Plaque Interactive Forces


Dental plaque can develop and maintain its position on the tooth surface only if the
interactive forces between the organism and the tooth surface are sufficiently strong
to withstand oral shear forces. Using an adhesion force spectroscopy approach, van
Hoogmoed et al. (2006) showed that Streptococcus mitis BMS biosurfactant changes
the interactive forces between mutant streptococci and enamel, explaining the effects of
biosurfactant on adhesion. This study showed for the first time direct measurements of
interactive forces between enamel and oral bacteria. The AFM probes were functional-
ized by gluing enamel particles onto silicon nitride cantilevers. Upon approach of the
enamel particles toward each of the mutant streptococci strains trapped in an isopore
polycarbonate membrane filter, force-distance curves were acquired, and the data are
presented in Table 2C.2.5. Microscopy

2C.2.11
Current Protocols in Microbiology Supplement 8
Figure 2C.2.7 At higher magnification, the honeycomb appearance of the surface lattice of Parame-
cium bursaria chlorella virus type 1 (PBCV-1) is evident. Scars and defects are common on the virion
faces, and the unique pentagonal clusters at the vertices can be seen in some cases. In panel D a large
fragment of a trisymmetronis flattened on the mica substrate, thereby allowing high-resolution imaging.
In panels E and F are two higher-magnification images of the network of trimeric proteins making up
the honeycomb arrangement. Reprinted from Journal of Structural Biology, vol. 149, Kuznetsov, Y.G.,
Gurnon, J.R., Van Etten, J.L., and McPherson, A. Atomic force microscopy investigation of a chlorella
virus, PBCV-1. pp. 256-263. Copyright 2005, with permission from Elsevier.

Figure 2C.2.8 AFM height image (6 × 6 µm, z range 1.5 µm), in aqueous solution, of a single
Saccharomyces cerevisiae cell trapped in a porous membrane. The cell (c) is located at the center
of the image, while the surrounding flat area represents the polymer membrane (m). The ∼1 µm
circular protrusion is attributed to a bud scar. Note that the cell is surrounded by an artifactual
structure (a) resulting from the contact between the AFM probe and the pore edges Reproduced
from Dufrêne (2002) with permission from the American Society for Microbiology.

Atomic Force
Microscopy

2C.2.12
Supplement 8 Current Protocols in Microbiology
Figure 2C.2.9 Mapping cell surface charges using chemically functionalized probes. (A) Force-distance curves
recorded in solutions of varying pH between the surface of Saccharomyces cerevisiae and an AFM probe function-
alized with carboxyl groups. (B) Adhesion force maps recorded at pH 7 and pH 4. The differences in adhesion forces
observed with pH were related to a change of ionization state of the cell surface. Abbreviation: nN, nanoNewton.
Reprinted from Current Opinions in Microbiology, vol. 6. Dufrêne, Y.F. Recent progress in the application of atomic
force microscopy imaging and force spectroscopy to microbiology. pp. 317-323. Copyright 2003, with permission
from Elsevier.

Investigating the Elastic Modulus of the Methanospirillum hungatei GP1 sheath


For the first time, AFM force spectroscopy measurements have allowed for probing
mechanical properties of microbial cell-surface layers on a local scale. In a pioneering
investigation, Xu et al. (1996) measured the elastic modulus of the sheath of the archeon
Methanospirillum hungatei GP1. The large values obtained for the Young modulus of
elasticity showed that this layer of unusual strength could withstand an internal pressure
of 400 atmospheres. Other investigators applied AFM to measure the local elasticity of
various cell-wall layers (Dufrêne, 2002). Touhami et al. (2003) used force mapping to
demonstrate that the cell wall elasticity of the S. cerevisiae varies significantly across the
cell surface. A 10-fold increase in the Young modulus of elasticity was measured for the
bud scar, in agreement with the accumulation of chitin in this region of the cell wall.

CONCLUSION
The AFM is the only instrument capable of nondestructively imaging specific molecules
in real time with a resolution comparable to electron microcopy. The exploration of
biological applications of the AFM, although technically challenging, are destined to
be of great importance. The new approaches that integrate AFM with classical optical
methods lead towards a deeper understanding of the dynamics of biological processes,
allowing a wide range of experiments, from discerning single molecule interactions
to monitoring the live cell responses. In this respect, AFM can be further combined
Microscopy
with optical methods to cover a much larger range of applications, including Forster
2C.2.13
Current Protocols in Microbiology Supplement 8
Table 2C.2.5 Characteristics of Force-Distance Curves Measured by AFM between Two Strains of Mutants Streptococci
and Enamel Particles with and without a Salivary Pellicle in the Absence and Presence of a Biosurfactant Coatinga,b

Approach curve characteristics Retraction curve characteristics


Bacteria Enamel coating F0 (nN) c
(nm) d
Adhesion (%)e Dmax (nm)f Fmax (nN)g

S. sobrinus 3.9 ± 1.7 53 ± 45 77 80 ± 78 −0.9 ± 0.9


HG 1025 Biosurfactant 14.3 ± 9.5 253 ± 110 0
Saliva 4.6 ± 1.4 83 ± 14 10 107 ± 44 −0.3 ± 0.2
Saliva + 4.3 ± 1.4 315 ± 90 8 178 ± 89 −2.0 ± 0.4
biosurfactant
S. mutans 0.9 ± 0.7 30 ± 11 78 39 ± 19 −0.8 ± 0.7
ATCC 25175 Biosurfactant 0.8 ± 0.4 59 ± 20 10 106 ± 110 −0.2 ± 0.1
Saliva 5.5 ± 1.2 186 ± 62 0
Saliva + 1.3 ± 0.8 333 ± 58 0
biosurfactant
a Results are means of 50 force-distance curves, taken over five different organisms on ten randomly selected locations per organism, and including
four different enamel particles standard ± deviations.
b Reproduced from van Hoogmoed et al. (2006) with permission of the International and American Associations for Dental Research.
c F the repulsive force of zero separation distance.
0
d , the decay length of the repulsive force upon approach.
e Percentage of the force-distance curves showing adhesion.
fD
max , average maximum distance at which a local maximum in adhesion forces occurred upon retraction.
gF
max , average local maximum in adhesion forces upon retraction.

resonance energy transfer (FRET; Ebenstein et al., 2004), surface plasmon resonance
(SPR) measurements (Chen et al., 1996), total internal reflection fluorescence (TIRF)
microscopy and/or internal reflection microscopy (IRM; Hugel et al., 2002; Trache and
Meininger, 2005), or confocal microscopy (Horton et al., 2000; Noy and Huser, 2003;
UNIT 2C.1).

The future should bring advances to the challenging aspects of the present technology,
most importantly the development of instrumentation for increasing acquisition speed to
allow for real time imaging of fast biological processes, an improved sample preparation
and immobilization technology, and better technology for preparing AFM functionalized
bioprobes.

LITERATURE CITED
Ahimou, F., Denis, F.A., Touhami, A., and Dufrêne, atomic force microscopy. Phys. Rev. E. 62:1034-
Y.F. 2002. Probing microbial cell surface 1044.
charges by atomic force microscopy. Langmuir Auerbach, I.D., Sorensen, C., Hansma, H.G., and
18:9937-9942. Holden, P.A. 2000. Physical morphology and
Amro, N.A., Kotra, L.P., Wadu-Mesthrige, K., surface properties of unsaturated Pseudomonas
Bulychev, A., Mobashery, S., and Liu, G.Y. putida biofilms. J. Bacteriol. 182;3809-3815.
2000. High-resolution atomic force microscopy Binnig, G., Quate, C.F., and Gerber, C.H. 1986.
studies of the Escherichia coli outer mem- Atomic force microscope. Phys. Rev. Lett.
brane: Structural basis for permeability. Lang- 56:930-933.
muir 16:2789-2796.
Binnig, G., Gerber, C.H., Stoll, E., Albrecht,
Arnoldi, M., Kacher, C.M., Bäuerlein, E., T.R., and Quate, C.F. 1987. Atomic resolution
Radmacher, M., and Fritz, M. 1998. Elastic with atomic force microscope. Europhys. Lett.
properties of the cell wall of Magnetospirillum 3:1281-1286.
gryphiswaldense investigated by atomic force Boonaert, C.J.P., Toniazzo, V., Mustin, C., Dufrêne,
microscopy Appl. Phys. A. 66:S613-S617. Y.F., and Rouxhet, P.G. 2002. Deformation of
Arnoldi, M., Fritz, M., Bäuerlein, E., Radmacher, Lactococcus lactis surface in atomic force mi-
Atomic Force
Microscopy M., Sackmann, E., and Boulbitch, A. 2000. croscopy study. Colloids Surf. B. Biointerfaces
Bacterial turgor pressure can be measured by 23:201-211.
2C.2.14
Supplement 8 Current Protocols in Microbiology
Boulbitch, A., Quinn, B., and Pink, D. 2000. Elas- croscopy/spectroscopy. Single Molecules 1:99-
ticity of the rod-shaped Gram-negative eubacte- 103.
ria. Phys. Rev. Lett. 85:5246-5249. Horton, M., Charras, G., Ballestrem, C., and
Butt, H.J., Downing, K.H., and Hansma, P.K. Lehenkary, P. 2000. Integration of atomic force
1990. Imaging the membrane protein bacteri- and confocal microscopy. Single Molecules
orhodopsin with the atomic force microscope. 1:135-137
Biophys. J. 58:1473-1480. Hugel, T., Holland, N.B., Cattani, A., Moroder, L.,
Camesano, T.A. and Logan, B.E. 2000. Prob- Seitz, M., and Gaub, H.E. 2002. Single molecule
ing bacterial electrosteric interactions using optomechanical cycle. Science 296:1103-
atomic force microscopy. Environ. Sci. Technol. 1106.
34:3354-3362. Karrasch, S., Hegerl, R., Hoh, J., Baumeister, W.,
Camesano, T.A., Natan, M.J., and Logan, B.E. and Engel, A. 1994. Atomic force microscopy
2000. Observation of changes in bacterial cell produces faithful high-resolution images of pro-
morphology using tapping mode atomic force tein surfaces in an aqueous environment. Proc.
microscopy. Langnuir 16:4563-4572. Natl. Acad. Sci. U.S.A. 91:836-838.
Chen, X., Davies, M.C., Roberts, C.J., Shakesheff, Kasas, S. and Ikai, A. 1995. A method for anchoring
K.M., Tendler, S.J. B., and Williams, P.M. 1996. round shaped cells for atomic force microscope
Dynamics surface events measured by simulta- imaging. Biophys. J. 68:1678-1680.
neous probe microscopy and surface plasmon Kuznetsov, Y.G., Malkin, A.J., Lucas, R.W., Plomp,
detection. Anal. Chem. 68:1451-1455. M., and McPherson, A. 2001. Imaging of viruses
Colton, R.J., Engel, A., Frommer, J.E., Gaub, by atomic force microscopy. J. Gen. Virol.
H.E., Gewirth, A.A., Guckenberger, R., Rabe, 82:2025-2034.
J., Heckel, W.M., and Parkinson, B. (eds.) 1998. Kuznetsov, Y.G., Gurnon, J.R., Van Etten, J.L., and
Procedures in Scanning Probe Microscopies. McPherson, A. 2005. Atomic force microscopy
John Wiley & Sons Ltd., Chichester, U.K. investigation of a chlorella virus, PBCV-1. J.
Costa, K.D. and Yin, F.C.P. 1999. Analysis of in- Struct. Biol. 149:256-263.
dentation: Implications for measuring mechani- Lal, R. and John, S.A. 1994. Biological applica-
cal properties with atomic force microscopy. J. tions of atomic force microscopy. Am. J Physiol.
Biomechan. Eng. 121:462-472. 266:C1-C21.
Dufrêne, Y.F. 2000. Direct characterization of the Lehenkari, P.P. and Horton, M.A. 1999. Single in-
physicochemical properties of fungal spores us- tegrin adhesion forces in intact cells measured
ing functionalized AFM probes. Biophys. J. by atomic force microscopy. Biochem. Biophys.
78:3286-3291. Res. Commun. 259:645-650.
Dufrêne, Y.F. 2002. Atomic force microscopy, a Lower, S.K., Tadanier, C.J., and Hochella, M.F.
powerful tool in microbiology. J. Bacteriol. 2000. Measuring interfacial and adhesion forces
184:5205-5213. between bacteria and mineral surfaces with
Dufrêne, Y.F. 2003. Recent progress in the appli- biological force microscopy. Geochim. Cos-
cation of atomic force microscopy imaging and mochim. Acta 60:2473-2480.
force spectroscopy to microbiology. Curr. Opin. Lower, S.K., Hochella, M.F., and Beveridge, T.J.
Microbiol. 6:317-323. 2001. Bacterial recognition of mineral surfaces:
Dufrêne, Y.F., Boonaert C.J.P., Gerin, P.A., Asther, Nanoscale interactions between Shewanella and
M., and Rouxhet, P.G. 1999. Direct probing -FeOOH. Science 292:1360-1363.
of the surface ultrastructure and molecular in- Meyer, G. and Amer, N.M. 1990. Simultaneous
teractions of dormant and germinating spores measurement of lateral and normal forces with
of Phanerochaete chrysosporium. J. Bacter. an optical-beam-deflection atomic force micro-
181:5350-5355. scope. Appl. Phys. Letters 57:2089-2091.
Ebenstein, Y., Mokari, T., and Banin, U. Möller, C., Allen, M., Elings, V., Engel, A., and
2004. Quantum-dot-functionalized scanning Müller, D.J. 1999. Tapping mode atomic force
probes for fluorescence-energy-transfer-based microscopy produces faithful high-resolution
microscopy. J. Phys. Chem. B 108:93-99. images of protein surfaces. Biophys. J. 77:1150-
Gad, M. and Ikai, A. 1995. Method for immobiliz- 1158.
ing microbial cells on gel surface for dynamic Morris, V.J., Kirby, A.R., and Gunning, A.P. (eds.)
AFM studies. Biophys. J. 69:2226-2233. 1999. Atomic Force Ficroscopy for Biologists.
Gad, M. and Ikai, A. 2001. Imaging living cells and Imperial College Press, London.
mapping their surface molecules with the atomic Müller, D.J. and Engel, A. 1999. Voltage and pH-
force microscope. In Methods in Cellular Imag- induced channel closure of porin OmpF visual-
ing (A. Periasamy, ed.) pp. 409-423, University ized by atomic force microscopy. J. Mol. Biol.
Press, Oxford. 285:1347-1351.
Hinterdorfer, P., Kienberger, F., Raab, A., Gruber, Müller, D.J., Baumeister, W., and Engel, A.
H.J., Baumgartner, G.W., Kada, G., Riener, C., 1996. Conformational change of the hexago-
Wielert-Badt, S., Borken, C., and Schlinder, nally packed intermediate layer of Deinococ-
H. 2000. Poly(ethylene glycol) an ideal cus radiodurans monitored by atomic force mi-
spacer for molecular recognition force mi- croscopy. J. Bacteriol. 178:3025-3030. Microscopy

2C.2.15
Current Protocols in Microbiology Supplement 8
Müller, D.J., Engel, A., and Amrein, M. 1997a. Sneddon, I.N. 1965. The relation between load
Preparation techniques for the observation of and penetration in the axisymmetric boussinesq
native biological systems with the atomic force problem for a punch of arbitrary profile. Int. J.
microscope. Biosens. Bioelectron. 12:867. Eng. Sci. 3:47-57.
Müller, D.J., Schoenenberger, C.A., Schabert, F., SPM Training Notebook. 2003. Veeco Instruments,
and Engel, A. 1997b. Structural changes in Santa Barbara, Calif.
native membrane proteins monitored at sub- Steele, A., Goddard, D.T., and Beech, I.B. 1994.
nanometer resolution with the atomic force An atomic force microscopy study of the biode-
microscope: A review. J. Struct. Biol. 119:149- terioration of stainless steel in the presence of
157. bacterial biofilms. Int. Biodet. Biodeg. 34:35-
Müller, D.J., Baumeister, W., and Engel, A. 1999. 46.
Controlled unzipping of a bacterial surface layer Tang, G., Yip, H.K., Samaranayake, L.P., Chan,
with atomic force microscopy. Proc. Natl. Acad. K.Y., Luo, G., and Fang, H.H.P. 2004. Direct
Sci. U.S.A. 96:13170-13174. detection of cell surface interactive forces of ses-
Müller, D.J., Heymann, J.B., Oesterhelt, F., Möller, sile, fimbriated and nonfimbriated Actinomyces
C., Gaub, H., Büldt, G., and Engel, A. 2000. spp. using atomic force microscopy. Arch. Oral
Atomic force microscopy of native purple mem- Biol. 49:727-738.
brane. Biochim. Biophys. Acta 1460:27-38. Touhami, A., Nysten, B., and Dufrêne, Y.F. 2003.
Noy, A. and Huser T.R. 2003. Combined force and Nanoscale mapping of the elasticity of micro-
photonic probe microscope with single molecule bial cells by atomic force microscopy. Langmuir
sensitivity. Rev. Sci. Instrum. 74:1217-1221. 19:4539-4544.
Ohnesorge, F., Heckl, W.M., Häberle, W., Pum, D., Trache, A. and Meininger, G.A. 2005. An atomic
Sára, M., Schindler, H., Schilcher, K., Kiener, force–multi optical imaging integrated micro-
A., Smith, D.P.E., Sleytr, U.B., and Binnig, scope for monitoring molecular dynamics in live
G. 1992. Scanning force microscopy studies cells. J. Biomed. Opt. 10:064023-064040.
of the S-layers from Bacillus coagulans E38– Trache, A., Trzeciakowski J.P., Gardiner L., Sun,
66, Bacillus sphaericus CCM2177, and of an Z. Muthuchamy, M., Guo, M., Yuan, S.Y., and
antibody binding process. Ultramicroscopy 42- Meininger, G.A. 2005. Histamine effects on en-
44:1236-1242. dothelial cell fibronectin interaction studied by
Ong, Y.-L., Razatos, A., Georgiou, G., and Sharma, atomic force microscopy. Biophys. J. 89:2888-
M.M. 1999. Adhesion forces between E. coli 2898.
bacteria and biomaterial surfaces. Langmuir Van Der Aa, B.C., Michel, R.M., Asther, M.,
15:2719-2725. Zamora, M.T., Rouxhet, P.G., and Dufrêne, Y.F.
Pum, D. and Sleytr, U.B. 1995. Monomolecular 2001. Stretching cell surface macromolecules
reassembly of a crystalline bacterial cell surface by atomic force microscopy. Langmuir 17:3116-
layer (S-layer) on untreated and modified silicon 3119.
surfaces. Supramol. Sci. 2:193-197. Van Der Mei, H.C., Busscher, H.J., Bos, R., de
Pum, D. and Sleytr, U.B. 1996. Molecular nano- Vries, J., Boonaert, C.J.P., and Dufrêne, Y.F.
technology and bioimetics with S-layers. In 2000. Direct probing by atomic force mi-
Crystalline Bacterial Cell Surface Proteins (U.B. croscopy of the cell surface softness of a fibril-
Sleytr, P. Messner, D. Pum, and M. Sara, eds.) lated and nonfibrillated oral streptococcal strain.
pp. 175-209. R.G. Landes, Austin, Texas. Biophys. J. 78:2668-2674.
Putman, C.A.J., Van Der Werf, K.O., De Grooth, Van Etten, J.L. (ed). 2000. Phycodna viridae Virus
B.G., Van Hulst, N.F., and Greve, J. 1994. Appl. Taxonomy, Classification, and Nomenclature of
Phys. Lett. 64:2454-2456. Viruses. Academic Press, London.
Radmacher, M., Tillman, R.W., Fritz, M., and Gaub, van Hoogmoed, C.G., Dijkstra R.J., van der Mei,
H.E. 1992. From molecules to cells: Imaging H.C., and Busscher, H.J. 2006. Influence of bio-
soft samples with the atomic force microscope. surfactant on interactive forces between mutans
Science 257:1900-1905. streptococci and enamel measured by atomic
Razatos, A., Ong, Y.-L., Sharma, M.M., and force microscopy. J. Dent. Res. 85:54-58.
Georgiou, G. 1998. Molecular determinants Wetzer, B., Pum, D., and Sleytr, U.B. 1997. S-
of bacterial adhesion monitored by atomic layer stabilized solid supported lipid bilayers.
force microscopy. Proc. Natl. Acad. Sci. U.S.A. J. Struct. Biol. 119:123-128.
95:11059-11064. Williams, T. 1998. Invertebrate iridescent viruses.
Schabert, F.A., Henn, C., and Engel, A. 1995. In The Insect Viruses (L.K. Miller and L.A. Ball,
Native Escherichia coli OmpF porin surfaces eds.) pp. 31-68. Plenum Press, New York.
probed by atomic force microscopy. Science Willemsen, O.H., Snel, M.M., Cambi, A., Greve,
268:92-94. J., De Grooth, B.G., and Figdor, C.G. 2000.
Scheuring, S., Ringler, P., Borgnia, M., Stahlberg, Biomolecular interactions measured by atomic
H., Müller, D.J., Agre, P., and Engel, A. force microscopy. Biophys. J. 79:3267-3281.
1999. High-resolution AFM topographs of the Wojcikiewicz, E.P., Zhang, X., and Moy, V.T. 2004.
Escherichia coli water channel aquaporin Z. Force and compliance measurements on living
Atomic Force EMBO J. 18:4981-4987.
Microscopy

2C.2.16
Supplement 8 Current Protocols in Microbiology
cells using atomic force microscopy (AFM). Dufrêne, 2002. See above.
Bio. Proceed Online 6:1-9. Dufrêne, 2003. See above.
Xu, W., Mulhern, P.J., Blackford, B.L., Jericho, The two references above are microbiology review
M.H., Firtel, M., and Beveridge, T.J. 1996. Mod- papers.
eling and measuring the elastic properties of an
Hansma, H.G. and Hoh, J.H. 1994. Biomolecu-
archaeal surface, the sheath of Methanospiril-
lar imaging with the atomic force microscope.
lum hungatei, and the implication of methane
Annu. Rev. Biophys. Struct. 23:115-139.
production. J. Bacteriol. 178:3106-3112.
Arnsdorf, M.F. and Xu, S. 1996. Atomic (scanning)
Yao, X., Jericho, M., Pink, D., and Beveridge, T.
force microscopy in cardiovascular research. J.
1999. Thickness and elasticity of Gram-negative
Cardiovasc. Electrophysiol. 7:639-652.
murein sacculi measured by atomic force mi-
croscopy. J. Bacteriol. 181:6865-6875. Giessibl, F.J. 2003. Advances in atomic force mi-
croscopy. Rev. Mod. Phys. 75:949-983.
Ikai, A. and Afrin, R. 2003. Toward mechanical
Key References manipulations of cell membranes and membrane
Morris et al., 1999. See above. proteins using an atomic force microscope. Cell
Covers AFM imaging on biological samples. Biochem. Biophys. 39:257-277.
Braga, P.C. and Ricci, D. (eds.) 2004. Atomic The four references above are review papers.
force microscopy: Biomedical methods and Heinz, W.F. and Hoh, J.H. 2005. Getting physical
applications. In Methods in Molecular Biology, with your chemistry: Mechanically investigating
Vol.242, Humana Press, Totowa, N.J. local structure and properties of surfaces with
Covers AFM imaging on biological samples and the atomic force microscope. J. Chem. Educ.
some functional studies. 82:695-703.
Colton, R.J., Engel, A., Frommer, J.E., Gaub, Covers AFM imaging and force spectroscopy—
H.E., Gewirth, A.A., Guckenberger, R., Rabe, theory and instrumentation.
J., Heckl, W.M., and Parkinson, B. 1998. Pro-
cedures in Scanning Probe Microscopies. Wiley Internet Resource
and Sons, West Sussex, England, U.K.. http://www.pacificnanotech.com/afm-
Good source of practical information on AFM artifacts single.html
imaging. Good presentation of imaging artifacts.

Microscopy

2C.2.17
Current Protocols in Microbiology Supplement 8

You might also like