Perspective Single-Molecule Biology: What Is It and How Does It Work?

Molecular Cell 24, 317–329, November 3, 2006 ª2006 Elsevier Inc.
DOI 10.1016/j.molcel.2006.10.017
Single-Molecule Biology:
What Is It and How Does It Work?
Jordanka Zlatanova1,* and Kensal van Holde2
1
Department of Molecular Biology
University of Wyoming
Laramie, Wyoming 82071
2
Department of Biochemistry and Biophysics
Oregon State University
Corvallis, Oregon 97331
Biochemistry and structural biology are undergoing
a dramatic revolution. Until now, we have tried to study
subtle and complex biological processes by crude in
vitro techniques, looking at average behaviors of
vast numbers of molecules under conditions usually
remote from those existing in the cell. Researchers
have realized the limitations of this approach, but
none other has been available. Now, we can not only
observe the nuances of the behaviors of individual
molecules but prod and probe them as well. Perhaps
most important is the emerging ability to carry out
such observations and manipulations within the living
cell. The long-awaited leap to an in vivo biochemistry
is at last underway.
Single-Molecule Approaches: General Introduction
Recent years have witnessed the emergence of an
entirely new field of science, often referred to as
single-molecule (SM) biology (or biochemistry) (Science,
1999; Journal of Biological Chemistry, 1999; Leuba and
Zlatanova, 2001; Chapter 16 in van Holde et al. [2006]).
SM methods represent a truly novel approach to biochemical/biological problems. All classical structural
and biochemistry/biophysics methods describe the behavior of enormous ensembles of molecules, averaging
the measured parameters over the entire molecular population. How any one molecule may behave over time
cannot be revealed by such studies; neither can the
behavior of individual molecules having different conformations and properties. It is important to realize that
seemingly homogenous populations of macromolecules that have no chemical differences do possess
intermolecular variations (often referred to as ‘‘static
disorder,’’ Kelley et al. [2001]). SM methods provide
the only available way to study their functional differences, by recording the behavior of individual members
of a certain population of molecules. In addition, SM approaches reveal fluctuations in the observable parameters of a single molecule over time, often with very high
temporal resolution, usually on the order of milliseconds. Assessing this ‘‘dynamic disorder’’ (Kelley et al.,
2001) is beyond the capabilities of population methods.
SM methodology provides a powerful tool for measuring important kinetic parameters. For example, if a macromolecule exists in two states and a SM signal, such as
fluorescence, can distinguish between them, one can
follow the behavior of the molecule in real time and determine a distribution of ‘‘dwell times’’ for each state.
The average dwell time t in a given state can simply be
*Correspondence: [email protected]
Perspective
determined by constructing a histogram of the measured dwell times and fitting its envelope to a single
exponential function. This should work for a simple (random) process, for which the dwell times should follow
a Poisson distribution (see Colquhoun and Hawkes
[1995]). In cases in which a single exponential fit does
not work, the process is either more complex than two
state or may involve some kind of memory effect (deviates from randomness). The dwell times of both states
can be obtained in a single experiment, without perturbing the system (which is not possible in bulk experiments), and can be converted into the familiar first-order
rate constants (the value of t in each state is the inverse
of the rate constant for the process that leads to transition to the alternative state). The SM methods do have,
at present, limitations in determining what is happening
during the change between the two states: this is simply
due to present limits of time resolution. In other words,
we are able to determine how long a molecule waits to
do something, but not how it does it. A major advance
in time resolution at the SM level would open a whole
new universe of understanding.
Another major and unique advantage that SM methodology provides is that it circumvents the need for synchronization of large numbers of molecules at a certain
stage of a biochemical process. Such synchronization
is difficult, often impossible, to attain. Even if the molecular population can be synchronized (usually by blocking all molecules at a certain stage of a process by, for
example, using drugs or withdrawing certain components from the mixture or by perturbing physical parameters), the release from the block quickly leads to loss of
synchrony because of the intrinsically stochastic nature
of individual reaction steps and the multistep character
of biochemical pathways.
SM methodology relies on the use of two general
types of approaches: those that allow observation of
single molecules under thermodynamic equilibrium or
nonequilibrium conditions (without externally imposed
perturbations of the system under study), and those
that study molecular behavior under applied force.
Examples of the first type of approaches include microscopic observation of fluorescently labeled single molecules and the so-called tethered-particle motion (TPM)
method, introduced in J. Gelles’s laboratory to follow
motions of a particle attached to a macromolecule,
which in turn is tethered to a surface (Schafer et al.,
1991; Yin et al., 1994). The second class of approaches
provides unique opportunities to investigate the mechanical responses of biological systems to applied tension and/or torsion (Khan and Sheetz, 1997; Leckband
and Israelachvili, 2001; for a brief overview, see Zlatanova and Leuba [2003a]). Forces can be applied by the
use of bendable beams (atomic force microscope
[AFM], optical or glass fibers) or by the so-called external-field manipulators (optical tweezers [OT], magnetic
tweezers [MT]) (Bustamante et al., 2000a). We begin
with a brief description of the principles of action of
the four most widely used SM methods. This will be followed by selected examples of how these methods can
Molecular Cell
318
be applied to study the mechanical properties of macromolecules or their complexes, the forces that govern
intra- and intermolecular interactions, and the functioning of biomolecules in various biochemical processes.
Finally, we will review some applications of these
methods to macromolecules in cells during their functioning in an in vivo context.
Physical Principles of SM Methods
Fluorescence
Fluorescence methods have been used for biological research for decades. Recent technical advances have
extended the use of fluorescent tags attached to macromolecules to the realm of SM biology. To that end, a single dye molecule is attached to a specific location in the
molecule of interest and then followed in a number of different ways to report on the behavior of that molecule.
Alternatively, fluorescent protein probes, like the green
fluorescent protein (GFP) or its variants (Tsien, 1998),
can be attached through recombinant DNA technology
to any protein of interest and used as fluorescent tags
in vitro or in vivo. The observables in SM fluorescence
studies include dye localization, fluorescence quenching by a nearby nonfluorescent quencher, fluorescence
anisotropy, and, most frequently, fluorescence (Fo¨rster)
resonance energy transfer (FRET) between two fluorophores. The theory and instrumentation behind these
methods have been reviewed in numerous publications
(Weiss, 1999, 2000; Lilley and Wilson, 2000; Selvin, 2000;
Ha, 2001a, 2001b; Deniz et al., 2001; Axelrod, 2001;
Michalet et al., 2003; Peterman et al., 2004; Neuweiler
and Sauer, 2004; Haustein and Schwille, 2003, 2004);
for some more application-oriented reviews, the reader
is also referred to Ha (2004) and McKinney et al. (2004).
A brief statement concerning criteria for fluorophores
is warranted here. They must (1) be bright (have high extinction coefficients and high quantum yield, i.e., be able
to transform a large proportion of the incident energy
into emitted light), (2) absorb and emit light in the visible
region of the spectrum, (3) show little fluctuation in the
emission intensity for the duration of the experiment,
(4) be relatively small so as not to perturb the molecule
under investigation, (5) be available in a form suitable
for covalent conjugation to the molecule, and (6) be photostable (Ha, 2001b). After w105 rounds of excitation
and emission, dyes lose their ability to be further excited, i.e., they undergo photobleaching. The lifetime
of a fluorophore strongly depends on the intensity of excitation, so if a prolonged (more than 5–10 s) period of
observation is required, it is advisable to reduce the excitation power to the possible minimum or to intermittently switch the incident light on and off.
Another property that needs to be considered in interpreting SM fluorescence data is the ‘‘blinking’’ behavior,
spontaneous transitions between bright (emitting) and
dark (nonemitting) periods (Ambrose et al., 1994; Xie
and Dunn, 1994) that occur before photobleaching.
The exact events causing blinking are unclear, but the
possibility of blinking should be taken into account
when interpreting spFRET data (see Sabanayagam
et al. [2005]).
The fluorescence SM detection methods fall into two
broad categories: wide-field and confocal. The widefield methods follow the behavior of immobilized mole-
cules over time by using two-dimensional detectors
such as CCD cameras. The biggest advantage of these
methods is the simultaneous observation of hundreds
of molecules in parallel, allowing the investigation of
both reversible and irreversible reactions and the detection of rare events. In the confocal method, the excitation beam is focused to a diffraction-limited spot
through an objective, which is also used to collect the
emitted fluorescence. The out-of-focus background
fluorescence is removed with a pinhole that is positioned at the back aperture of the objective. The detection of the fluorescent spot is done with photomultiplier
tubes or avalanche photodiodes (APDs) (for specific usage of these detectors, see Ha [2001b]). The confocal
setups allow observation of surface-immobilized molecules, as well as of molecules traversing (by diffusion
or flow) the small volume of liquid excited by the laser.
Confocal detection is also used in a couple of related
fluorescence techniques that study the behavior of
a small number of molecules (fluorescence correlation
spectroscopy and two-color crosscorrelation spectroscopy, see Haustein and Schwille [2003, 2004]).
Wide-field instrumentation uses two different approaches for excitation: epifluorescence and total internal reflection (TIR). In standard epifluorescence, the light
is sent through the epi-illumination port of conventional
fluorescence microscopes; thus, both the microscope
optics and the sample contribute to out-of-focus background fluorescence, reducing the signal-to-noise ratio.
In TIR, the excitation light is directed toward an interface
between two media of different refractive indices (i.e.,
from an optically denser medium, such as a glass slide,
into a less-dense medium, such as water) (Axelrod,
2001). The incident angle of the beam is set larger than
a certain critical angle, defined by the properties of the
two media; all the light is reflected off the glass and
does not penetrate into the solution (Figure 1A). However, an electromagnetic field that oscillates with the
same frequency as the incident light does form in the
less-dense medium (the water on the other side). Because this electromagnetic field (also called evanescent
field or wave) decays exponentially from the glass surface, it is capable of exciting fluorophores only in
a very small volume close to the surface, thus effectively
preventing out-of-focus fluorescence background. The
excitation light itself is cleanly removed from the observation chamber, reducing the background even further.
The two types of TIR fluorescence microscopy—
prism-based and objective-based—are described in
an excellent recent tutorial (Knight et al., 2005) (see
also Peterman et al. [2004]; Zheng et al. [2005]).
Of all SM fluorescence methods, we will briefly introduce only FRET, as this method is being most widely
used. FRET relies on nonradiative transfer of electronic
excitation energy between two fluorophores, a donor
and an acceptor. The theory developed by Fo¨rster
(1959) predicts that the efficiency of energy transfer is
a function of the distance (R) between the two dyes:
EFRET = 1/[1 + (R/Ro)6], where Ro is the interdye distance
at which half of the energy is transferred. Ro depends on
a number of characteristics of the dye pair (see Selvin
[2000]; Ha [2001b]; Neuweiler and Sauer [2004]). Following the experimental verification of the theory (Stryer
and Haugland, 1967), EFRET has been widely used as
Perspective
319
Figure 1. Detecting Fluorescence from Individual Fluorophores by
Total Internal Reflection Fluorescence Microscopy and Principle of
Single-Pair FRET for Studying Biological Macromolecules
(A) In total internal reflection fluorescence (TIRF) microscopes, the
sample is excited by the intense electromagnetic field that is created
by the totally reflected laser beam at the interface of two media of
different refractive indexes (usually quartz and water). This field,
also known as an evanescent wave, decays exponentially with the
distance from the interface and thus excites only the fluorophores
present in a small volume in close proximity to the quartz. The figure
illustrates the more commonly used prism-based setup; the advantages of the alternative, objective-type TIRFM are discussed in
Knight et al. (2005).
(B) Principle of spFRET (see text).
a spectroscopic ruler. Because Ro for many dye pairs
lies in the vicinity of 5–6 nm, FRET is best used to measure distances in the range of 2–8 nm (see illustration of
principle in Figure 1B). This distance range is not
covered by any other solution technique. One should
keep in mind that the distances determined are not
to be taken as absolute values but rather used for
comparative purposes because Ro may depend on
dye environment.
Since measuring FRET requires recording of the donor and acceptor emissions in different channels, dye
pairs for spFRET should have a large spectral separation
so as to avoid leakage of the donor signal into the acceptor channel and direct excitation of the acceptor by
the laser. It is also highly desirable to choose pairs
with comparable emission quantum yields so that the
occurrence of FRET is clearly recordable as anticorrelated changes in the fluorescence intensities of both
dyes: if the intensity of the donor emission falls, the intensity of the acceptor emission should simultaneously
increase, and vice versa. Finally, one has to be aware
of blinking effects (see above). Dark periods of the donor
do not pose problems in interpretations, since no
fluorescent signal is available during these periods. If
the acceptor, on the other hand, blinks, a concomitant
increase in donor emission will be observed and possibly misinterpreted as a loss of FRET due to conformational changes or biochemical processes. Since the
emission ‘‘on’’ time is dependent on the excitation intensity, and the efficiency of FRET is not, blinking can be
distinguished from actual distance changes in the molecules by varying the intensity of the exciting radiation.
Atomic Force Microscope
The AFM (Binnig et al., 1986) has opened new horizons
in both imaging and force measurements. In the AFM,
a sharp probe (tip) is mounted at the end of a flexible
cantilever and allowed to interact with a surface-immobilized sample. Precise lateral and vertical displacement
of the sample with respect to the probe is achieved by
a computer-controlled piezoceramics stage holding
the sample (Figure 2A) or, conversely, the cantilever
holder. Interaction forces between atoms on the sample
and atoms on the tip cause deflection of the cantilever
that is registered by a laser reflected off its back into
a photodiode detector. These deflections can be converted into a topographic image of the sample (when
the probe is raster scanned in the x-y plane) or used to
produce force-distance curves (when the probe is
moved in the z direction only). AFM imaging has become
routine and has led to significant advances in our understanding of how biological molecules function. Although
the resolution of AFM is comparable (for soft biological
samples) to that of the electron microscope, AFM has
the enormous advantage in that it can deal with moist
or even buffer-immersed samples, allowing imaging under physiologically relevant conditions.
The application of AFM for force spectroscopy has
also significantly contributed to our understanding of
intra- and intermolecular interaction forces, especially in
the range of 50 pN to 1–2 nN (for some earlier examples,
see Table 1 in Zlatanova et al. [2000]). Force-distance
curves are obtained upon moving the probe upwards
(retraction curve) after it has been pushed into the sample (approach curve); the signature of the retraction
curve can be used to measure interactive forces (a
more detailed description of the types of interaction
forces that are ‘‘felt’’ by the cantilever during the approach and retraction portions of the force curve can
be found in Zlatanova et al. [2000]). An illustration of
a force curve obtained upon stretching of proteins is
presented in Figure 2B.
Optical Tweezers
The OT technique has been extensively reviewed both in
terms of theory and instrumentation (e.g., Visscher and
Block, 1998; Smith et al., 2003; Leuba et al., 2003; Bennink et al., 2005; Knight et al., 2005) (for a fascinating account of the history of the development of optical trapping, see Ashkin [1997]). Light can exert forces on
small transparent beads in such a way that the bead is
kept suspended at a point close to the waist of a focused
laser beam (Figure 3A). If such a ‘‘trapped’’ bead is
pulled out of its equilibrium position by whatever external force, a net restoring force resulting from the bead’s
interaction with light will effectively pull it back toward
the equilibrium position. To a first approximation, the
trap can be considered as a Hookean spring, with the
force determined by F = kDx, where k is the spring
Molecular Cell
320
examples of using OT are presented below (Figures
3C, 3D, 5C, 5D, 6C, and 6D).
Magnetic Tweezers
MT is another force-field manipulation technique in
which the macromolecule is acted upon from a distance,
this time through a controllable magnetic field (Figure 4A) (Strick et al., 1996, 2000; Zlatanova and Leuba,
2003b; Leuba et al., 2003; Revyakin et al., 2005). The
macromolecule is attached between a surface and
a superparamagnetic bead. Manipulation of the external
magnetic field allows application of stretching forces to
the tethered molecule, the magnitude of the force being
precisely controlled by the distance between the cuvette and the external magnet(s). The force on the
bead, and thus the tether, is measured using the equipartition theorem, with two observables: the extent of
Brownian motion in the x-y plane and the tether length.
A major advantage of this instrument is the ability to
introduce controlled levels of torsional stress in the molecule by simply rotating the magnetic field clockwise or
counterclockwise, depending on the sign of supercoiling one wants to achieve. Due to space limitations, we
have only illustrated this powerful technique in Figure 4B, but the reader is referred to the recent review
by Charvin et al. (2005) that focuses on the application
of MT to the study of topoisomerases; Zlatanova et al.
(2006) have discussed the application of MT (among
other methods) to the study of transcription initiation
and elongation.
Figure 2. Use of the Atomic Force Microscope for Imaging and
Force Spectroscopy
(A) Schematic of the principle of action of the AFM (see text).
(B) A schematic representation showing the structural transitions in
a multidomain protein upon mechanical stretching using the AFM.
The sawtooth pattern of the force-extension curves results from
strings of successive enthalpic and entropic portions, reflecting
the unfolding of individual domains, followed by entropic stretching
of the unfolded domain. The unfolding of each domain adds significant length to the chain and relaxes the stress on the cantilever,
which returns to its nondeflected state. The unfolded portion of
the polypeptide chain can now undergo entropic stretching, which
is accompanied by gradual deflection of the cantilever (adapted
from Zlatanova et al. [2000]).
constant of the trap and Dx is the displacement of the
bead from the focus of the trap.
In OT applications, the molecule is tethered between
the optically trapped bead and a movable platform (another bead, usually held in a micropipette, or a cover
slip) (Figure 3B). The platform can be moved in a controllable way to apply tension (and, recently, torsion) to the
tether to record force-extension curves. Some specific
Applications of SM Methodology to Biological
Problems
Mechanical Properties of Biological Macromolecules
and Their Complexes
A conceptual framework for these kinds of studies has
been provided by Bustamante et al. (2004). DNA was
the first biomacromolecule to be investigated under
applied mechanical force by a variety of SM methods
(Smith et al., 1992, 1996; reviewed in Bustamante et al.
[2000b, 2003]; Zlatanova and Leuba, 2002, 2003a,
2004). An example stretching-relaxation curve obtained
with OT is presented in Figure 3C. The stretching curve
reveals different behavior of the molecule under different force regimes: (1) up to 10 pN, the molecule behaves
as a flexible polymer, accurately described by the wormlike chain model; (2) above 10 pN, the chain has been extended, and the DNA now behaves as a stretchable
solid, lengthening beyond its B form contour length;
and (3) at forces exceeding 65 pN, the molecule suddenly yields, changing conformation to an overstretched
form (w1.7 times its contour length). The nature of
the overstretching transition, frequently referred to as
B-to-S (stretched) transition, is still debated.
OT have been also used to stretch individual chromatin fibers or nucleosomal arrays, either directly reconstituted in the flow cell of the instrument (Bennink et al.,
2001; Figure 4D) or preassembled from defined arrays
of nucleosome positioning sequences and purified
core histones (Brower-Toland et al., 2002). The high
temporal and spatial resolution achieved in these experiments led to defining the force (20–40 pN) needed to unravel the DNA from individual nucleosomal particles (for
further discussion, see Zlatanova and Leuba [2002,
2003a, 2004]; Zlatanova [2003]).
Perspective
321
Figure 3. Optical Trapping and Its Application for Stretching DNA and Chromatin Fibers
(A) The behavior of a dielectric bead in a focused laser beam (see text).
(B) OT can be used to study macromolecules by attaching them to the optically trapped bead at one end and to a movable platform (in the illustration, another bead held in a micropipette) at the other end. Moving the platform away from the trapped bead will apply stretching forces
to the macromolecule, causing displacement of the trapped bead from its equilibrium position. This displacement can be measured and used to
estimate the forces applied to the molecule.
(C) Stretching of l-DNA. A typical DNA force-extension curve exhibits three portions: entropic stretching of the worm-like chain at the beginning,
enthalpic (elastic) stretching of the structure at lengths of the molecule approaching or slightly exceeding the contour length (in this case
w16 mm), and finally, at w65 pN force, a transition from the B to the so-called S (stretched) form. The physical nature of the S form has been
controversial (for detailed discussion, see Bustamante et al. [2000b, 2003]; Zlatanova and Leuba [2002, 2003a, 2004]). The DNA was stretched
to w20 mm (red curve) and then relaxed (blue curve).
(D) Stretching of a chromatin fiber assembled on naked l-DNA molecule by the addition of X. laevis egg extract directly into the flow cell of the
instrument. The extract contains core histones and protein factors needed for assembly (assembly is manifested by shortening of the distance
between the two beads with time). Note the sharp discontinuities in the force-extension curve reflecting the unraveling of the DNA from around
the histone octamer that forms the core of the nucleosomal particles. Nucleosomes can unwrap either individually or in groups of two, three, or
four. At high extension, when all histones have been forced off, the curve approaches that of naked DNA. Note that the two force-extension
curves for DNA (C) and chromatin (D) are aligned with respect to the length of the structure during stretching, so that a direct comparison of
the behavior of DNA and of chromatin is possible.
(C) and (D) are modified with author permission (Bennink, 2001).
Protein unfolding as a result of mechanical stretching
has been one of the major applications of AFM. The
groundbreaking papers of Rief et al. (1997) and Oberhauser et al. (1998) showed that unfolding of proteins
that contain numerous individually folded domains produces a sawtooth pattern in the force-extension curves
reflecting successive unfolding of each individual domain (the appearance of a typical force-extension curve
and its interpretation are illustrated on the example of
a recombinant polyprotein consisting of tandemly repeated monomeric domains; see Figure 2B). A series
of publications, mainly from J. Fernandez’s laboratory,
revealed (1) the existence of unfolding intermediates in
which each domain abruptly extends by a certain length
before the actual full unfolding (Marszalek et al., 1999a),
(2) the occurrence of misfolding events that can be detected upon repeated extension/relaxation cycles
(Oberhauser et al., 1999), (3) domain length polymorphism at a resolution comparable to that of structural
methods, such as NMR and X-ray crystallography (Carrion-Vazquez et al., 1999), and (4) folding trajectories of
single proteins upon force relaxation (Fernandez and
Li, 2004; note that the interpretations of this work have
been questioned, Sosnick [2004]; Best and Hummer
[2005]). An interesting technical development allowed
the use of AFM to calibrate the distance-dependent
decay of an evanescent wave (see above); simultaneous
measurements during stretching of fluorescence
Molecular Cell
322
Figure 4. Magnetic Tweezers and Their Application for Studying the
Response of DNA to Torsion
(A) Schematic of the principle of action of MT (see text).
(B) Relative extension of a DNA molecule as a function of the degree
of imposed supercoiling (reproduced from Strick et al. [1998], with
permission from the Biophysical Society; the figure has been slightly
modified in terms of the labeling of the lines). MT were used to introduce a controlled level of supercoiling density in a single molecule of
DNA tethered between a surface and a superparamagnetic bead in
a topologically constrained way. The changes in tether length (extension) were followed as the external magnets were rotated either
clockwise or counterclockwise, to create negative or positive supercoiling tension, respectively. Force-extension curves taken at different levels of stretching forces reveal intriguing differences in the behavior of positively versus negatively supercoiled DNA. The linear
shortening of the molecules results from plectoneme formation,
whereas the portions of the curves that remain parallel to the x
axis despite the continuous pumping in of superhelical tension reflect structural transitions in the molecules that are not accompanied by shortening. In the case of negative superhelical tension,
the helix undergoes local denaturation (Strick et al., 1998). In the
case of positive superhelical tension, the B-DNA undergoes a structural transition to a new phase called P-DNA (Pauling-DNA, Allemand et al. [1998]); in this conformation, the phosphate-sugar backbones of the two strands are wound inside the structure, with the
bases exposed to the solution.
intensity and protein length from a single ubiquitin polyprotein resolved w20 nm steps in the unfolding of ubiquitin (Sarkar et al., 2004). Finally, a recent paper identified a folding intermediate of individual ribonuclease
molecules stretched and relaxed with OT and defined
the energy landscape of this enzyme (Cecconi et al.,
2005).
A striking example of the resolution power of SM
methods to investigate the mechanical properties of biomacromolecules under tensile stress is the identification of chair-boat transitions in the glucopyranose ring
of polysaccharides as the basis for the molecule’s elasticity (Marszalek et al., 1998). Further, a two-step conversion from chair to boat and then to an inverted chair
has been reported; it was postulated that chair inversion
results from torque generated by force application to axial (but not equatorial) glycosidic bonds (Marszalek
et al., 1999b).
SM methods have also been used to study the behavior of double-helical DNA upon twisting. Croquette and
Bensimon pioneered the use of MT to investigate the
structural transitions of dsDNA under superhelical tension (Strick et al., 1996; Allemand et al., 1998; reviewed
in Strick et al. [2000]; Zlatanova and Leuba [2003b]).
These investigators topologically constrained a single
DNA molecule between a glass surface and a magnetic
bead by attaching all four ends of the molecule through
multiple points to the respective surfaces. Such an attachment prevented DNA from swiveling about its anchoring points, allowing the introduction of superhelical
tension in the molecule by precisely controlled rotation
of the external magnetic field. By recording the relative
extension of the molecule as a function of superhelical
density at fixed force values, differences in the behavior
of positively versus negatively supercoiled molecules
were revealed (Figure 4B) and structural transitions (denaturation or formation of Pauling-like DNA structures)
deduced.
Understanding Transcription
The wealth of information already available from biochemistry and high-resolution structural studies has
been recently supplemented by SM data. All stages of
transcription have been investigated, beginning with
the search for and binding to the promoter region, the
unwinding of the DNA double helix at the promoter,
the initial synthesis of abortive transcripts, the transition
to elongation, elongation itself, and termination. It is beyond the scope of this tutorial to cover the SM transcription field, but a comprehensive review has been recently
published (Zlatanova et al., 2006). Here, we will only illustrate how two methods, TPM and OT, have been
used to study elongation, providing unprecedented insights into the ‘‘static’’ heterogeneity of individual molecules in a RNAP population and into the different kinds
of pauses that intersperse the overall monotonic movement of the enzyme along the DNA template. Yin et al.
(1994) made use of the TPM method designed earlier
(Schafer et al., 1991) (Figure 5A) to obtain real-time trajectories of the longitudinal movement of the DNA template past the catalytic center of individual immobilized
RNAPs (Figure 5B). The statistical evaluation of such trajectories led them to conclude that the kinetic properties
of individual RNAP molecules in a putatively homogeneous population varied to a larger degree than that expected on the basis of the experimental uncertainty of
the rate measurements. Recently, the TPM method
was modified to avoid the use of immobilized enzymes
and thus a potential artifactual source of enzyme heterogeneity (Tolic´-Nørrelykke et al., 2004). The same
Perspective
323
Figure 5. Studying Transcription Elongation
(A) Schematic of the tethered particle motion
(TPM) method, based on the dependence of
the extent of Brownian motion of a bead
(readily measured by light microscopy) on
the length of the DNA template tethering it
to the surface. In transcription experiments,
the polymerase is immobilized on a glass surface and a bead is attached to either the upstream or the downstream end of the DNA
template; elongation will result in lengthening
or shortening of the tether (and a change in
the Brownian motion of the bead), depending
on the particular geometry used.
(B) Transcription elongation as measured by
the TPM method. Example tether-length/
time courses with the bead at the upstream
(left traces) or downstream (right traces) end
of the template (reproduced from Yin et al.
[1994] with permission from the Biophysical
Society).
(C) Concept of recent experiments from
S. Block’s laboratory using the novel passive,
all-optical force clamp that takes advantage
of the existence of a region in the trap where
the force is approximately constant for small
bead displacements (Greenleaf et al., 2005;
Herbert et al., 2006). Two beads are held in
separate optical traps: the transcribing polymerase is attached to one of the beads and the upstream end of DNA to the other bead, creating
a bead-DNA-RNAP-bead dumbbell geometry. In the configuration used, the trap applies an assisting load. This instrument achieves angstrom-level resolution.
(D) Representative records of transcription along a repetitive DNA template containing eight repeats of an w230 bp sequence that possesses
previously identified pause sites. Most records display distinct pauses at the expected pause sites; in addition, other pause sites are identified
that display sequence similarities (reprinted from Herbert et al. [2006] with permission from Cell Press).
conclusion was reached, recently confirmed by Herbert
et al. (2006), using OT.
Block’s group has recently developed a passive, alloptical force clamp OT (Greenleaf et al., 2005) that is
highly stable, allowing lateral resolution in the angstrom
range. Using a dumbbell geometry—bead-DNA-RNAPbead—the group has shown that the polymerase advances in steps of w3.4 A˚, i.e., a translocation step of
one base pair follows each catalytic step of adding
one base to the nascent transcript (Abbondanzieri
et al., 2005). Using a similar geometry (but rather different instrumentation) (Figure 5C), they have also extensively studied the nature of the pauses experienced by
the elongating enzyme. It was convincingly demonstrated that the short, frequent, ubiquitous pauses (Figure 5D) are sequence dependent (Herbert et al., 2006)
and, as earlier work has shown, are characterized by
sharp transitions into and out of the pause (Neuman
et al., 2003). During the long (>20 s) pauses, the RNAP
backtracks along the template with extrusion of an already synthesized portion of the RNA transcript from
the catalytic center (Shaevitz et al., 2003). Sophisticated
averaging procedures allowed probing details of polymerase motion that could have remained masked by
noise in records of individual RNAP molecules. Three
successive phases of the long pause were discerned:
backtracking at the beginning, pausing per se, and
slow downstream recovery. The duration and frequency
of the backtracking pauses are strongly dependent on
application of either opposing or assisting force, in contrast to the short pauses, which are load independent.
Understanding Molecular Motors Translocating
on Linear Tracks
Molecular motors are enzymes that use the energy from
NTP hydrolysis to translocate along tracks (DNA in the
case of polymerases, helicases, and nucleosome remodeling factors, etc., or actin filaments or microtubules
in the case of cargo transporters, such as myosin, kinesin, and dynein). SM studies of RNA polymerases have
been described above. We now consider as an example
the application of SM techniques to the study of kinesin
motion on microtubules.
There have been two categories of mutually excusive
models constructed on the basis of structural and biochemical studies. Each of the models has specific predictions on the movement of the two heads and on the
rotation of the stalk during translocation. The SM studies illustrated in Figure 6 show the variety of methodologies used and the evolution of our knowledge on the motion mechanism of kinesin (for details, see the figure
legend and the original papers cited there; reviewed in
Yildiz and Selvin [2005b]).
Gelles and coauthors performed a SM microtubule
gliding assay, which involved immobilization of the kinesin tail regions onto a glass surface; the translocation of
microtubules was followed by microscopy (Hua et al.,
2002; Figure 6A). No rotation of the microtubules was
observed (Figure 6B); the data were interpreted in terms
of inchworming motion of the kinesin along the microtubule but were also consistent with an asymmetric, handover-hand motion. Two groups, using slightly different approaches, observed ‘‘limping’’ of kinesin during
Molecular Cell
324
Figure 6. Representative SM Data on Kinesin Moving Along Microtubules
(A) Experimental geometry for observing stalk rotation: a microtubule is bound to a single-surface-attached kinesin molecule.
(B) In the presence of ATP, there is no rotation of the stalk, as seen by the overlaid traces (arrows) of a microtubule imaged at 1 s intervals. The
right- and the left-hand traces show the same microtubule in two different time periods separated by 104 s (adapted from Hua et al. [2002] and
reproduced with permission from AAAS).
(C) Time course of the displacement of a single heterodimeric kinesin molecule, as visualized by OT nanometry (three left traces). This molecule
contains two heads with very different mechanochemical cycle rates (wild-type and a single point mutant in the nucleotide-binding motif, with
much-reduced microtubule-gliding speed). The traces indicate that in most cases the steps are 16 nm, but sometimes a shoulder at 8 nm is distinguishable. Thus, the observed 16 nm step is actually two successive steps of very different dwell times. The wild-type homodimer (two right
traces) shows clear 8 nm steps of almost equal duration (reproduced from Kaseda et al. [2003] with permission from Nature Publishing Group,
McMillan Publishers Ltd.).
(D) Stepwise motion of recombinant kinesin homodimers as measured by OT: intrinsic stepping rate can alternate between two different values at
each sequential step, causing the molecule to limp. The construct tested was DmK401, a truncated derivative of D. melanogaster kinesin containing two identical heads and a sufficient length of the neck coiled coil for dimerization. The dwell times between successive 8 nm steps alternate
between slow and fast phases, causing steps to appear in pairs (red and blue). The dwell times in the record are numbered sequentially to make the
point that the longer-lived dwell times tend to cluster systematically, in this specific trace, in the odd-numbered subset. The vertical bars indicate
the occurrence of the steps (traces reproduced from Asbury et al. [2003], with permission from AAAS).
(E) A typical image of surface-immobilized Cy3-DNA acquired within 0.5 s using objective-type TIRF. The image contains w14,000 photons and
can be fitted to a two-dimensional Gaussian, which approximates the point spread function (PSF) of the Cy3; the center of the PSF yields w1 nm
precision and accuracy in the center localization of the fluorescence spot (reproduced from Yildiz and Selvin [2005a] with permission from the
American Chemical Society). This technique has been named FIONA (Yildiz et al., 2003; Yildiz and Selvin, 2005a).
(F) Position versus time for kinesin motility as determined by FIONA: example traces for an E215C homodimer kinesin. Dots, experimental points;
red lines show when steps occur (vertical jumps) and the average positions between steps (plateaus). Both the step sizes of an individual head of
a kinesin dimer and dwell-time analysis support a hand-over-hand mechanism (reproduced from Yildiz et al. [2004] with permission from AAAS).
Perspective
325
Table 1. Selected Examples of Single-Molecule Studies In Vivo
Object of Study and Technique
Main Observations/Conclusions
References
Lateral motion of fluorescently labeled lipids
inserted into membranes of living cells;
single-dye tracking by epifluorescence
microscopy
A saturated lipid probe showed confined
diffusion within small areas of w0.7 mm,
whereas an unsaturated lipid diffused freely in
the membrane; lipid domains (rafts) exist in cell
membranes as areas with high local
enrichment of saturated lipids
One molecule of EGF interacts with an EGFR dimer,
followed by binding of a second EGF molecule;
formation of dimers of receptor molecules is
directly visualized; EGFR becomes
phosphorylated after dimerization
The cAMP gradient in the medium (that causes
the chemotactic movements) is converted to
a spatial difference in the reaction state of the
cAMP receptor, with cAMP signaling more
active in the anterior region of the cell
The distribution of diffusion constants within
cells is broader than that obtained in glycerol
solution, from simulations or theory; the
intracellular motion cannot be described by
simple diffusion
Protein transport to the plasma membrane
occurs in phases: docking of the Golgi carrier
onto the membrane, its flattening, and
diffusion of the protein within the membrane;
both small spherical and large tubular carriers
participate in the process
Diffusion trajectories indicate complex motions,
including free and anomalous diffusion, and
directed motion by motor proteins both in the
cytoplasm and nucleus; much faster infection
than measured in bulk experiments
Unsaturated lipids are confined within 230 nm f
compartments for several milliseconds before
hopping to adjacent compartments;
the 230 nm compartments exist within larger
w750 nm domains where lipids are
confined for w0.33 s; diffusion in cellular
membranes is slow because of
compartmentalization
Polymerization and depolymerization occur
throughout lamellipodia with constant kinetics;
polymerization occurs away from the tip of the
lamellipodia
The movement (before viral fusion with
endosomes) occurs in three stages: (1) slow,
actin-dependent movement in the cell
periphery, (2) rapid unidirectional dyneindirected movement towards the nucleus, and
(3) intermittent, often bidirectional movement
on microtubules in the perinuclear region
The diffusion of the lipid probe is blocked in the
initial segment of the axon membrane by
transmembrane proteins anchored to actinbased meshwork under the membrane
Schutz et al., 2000
Epidermal growth factor receptor (EGFR)
signaling in membrane of living carcinoma cells;
single-molecule tracking by objective-type TIRF;
spFRET between Cy3-EGF and Cy5-EGF
Association and dissociation of individual
Cy3-cAMP molecules with the cell surface
of Dictyostelium during chemotaxis;
single-molecule tracking by objective-type TIRF
Intracellular motion of R-phycoerythrin
(autofluorescent protein containing
w30 chromophores); single-molecule
tracking by epifluorescence
Final stages of constitutive exocytosis
(fusion of single post-Golgi carriers with
the plasma membrane); single-molecule
tracking by objective-type TIRF
Viral infection by adeno-associated viral
particles; single viral particle tracking
by epifluorescence
Movement of unsaturated phospholipids in the
plasma membrane of rat kidney fibroblasts;
single-molecule tracking by objective-type TIRF;
single-particle tracking of a 40 nm gold bead
attached to an unsaturated lipid probe by
videomicroscopy
Polymerization and depolymerization of actin
filaments in lamellipodia; single-molecule
tracking of actin-GFP (SM speckle analysis)
by epifluorescence
Constitutive endocytosis of influenza virus;
Single viral particle tracking by epifluorescence
Distribution of membrane proteins in different
regions of polarized neuron membranes;
single-molecule tracking by objective-type
TIRF; single-particle tracking of a 40 nm
gold bead attached to an unsaturated
lipid probe by video-microscopy;
dragging experiments on the
gold-particle-labeled probe by OT
DNA uptake during bacterial transformation; OT
Intracellular signal transduction pathway
(activation of the oncoprotein Ras by epidermal
growth factor); spFRET by objective-type TIRF
DNA molecules are transported at a linear rate
without observable pausing, reversals, or
slipping; the uptake rate is force
independent to up to w40 pN of opposing
force; proton motive force is used for
DNA transport
Activation of single Ras molecules leads to
a considerable slowing down of their diffusion;
the suppressed diffusion of Ras suggests the
formation of large Ras-signaling complexes in
the membrane
Sako et al., 2000b
Sako et al., 2000a;
Ueda et al., 2001
Goulian and Simon, 2000
Schmoranzer et al., 2000
Seisenberger et al., 2001
Fujiwara et al., 2002
Watanabe and
Mitchison, 2002
Lakadamyali et al., 2003
Nakada et al., 2003
Maier et al., 2004
Murakoshi et al., 2004
(Continued on next page)
Molecular Cell
326
Table 1. Continued
Object of Study and Technique
Main Observations/Conclusions
References
Protein translocation through nuclear pore
complexes (NPC) by using a fluorescent model
substrate (two identical GFP domains are
linked together); single-molecule tracking
by epifluorescence
Protein molecules spend w10 ms within NPC,
moving rapidly in a random walk though the
central pore; the rate-limiting step is the
substrate exit from the pore; NPC must be
capable of transporting w10 molecules
simultaneously
Single quantum dots diffused at
characteristically different rates in different
regions of the membrane or inside the
cytoplasm
T cell activation causes coclustering of several
membrane proteins in discrete microdomains;
two-color imaging reveals that these
microdomains exclude or trap specific
proteins; diffusional trapping through proteinprotein interactions participate in T cell
signaling
Peroxisomes move in steps in both anterograde
(kinesin) and retrograde (dynein) directions,
with step size of w8 nm; each peroxisome
takes a few steps driven by dynein, then
a couple by kinesin, etc.; dynein and kinesis
work in coordination; several kinesins or
dyneins can work in concert to move
the cargo faster
Yang et al., 2004
Avidin-CD14 receptor movement in HeLa cells;
single-particle tracking of biotinylated
peptide-coated quantum dots by
epifluorescence
Signaling proteins in immortalized T cells upon
T cell activation; combination of scanning
confocal (population) methods and
single-molecule tracking by objective-type TIRF
Organelle (peroxisone) movement in cultured
Drosophila cells; FIONA
walking (Kaseda et al., 2003; Asbury et al., 2003; Figures
6C and 6D); such limping (alternating nonequivalent behavior of the two heads during successive translocation
steps) excluded both the inchworm type of motion and
a symmetric hand-over-hand motion. Finally, the fluorescence imaging with one nanometer accuracy (FIONA)
studies of Selvin’s group (Yildiz et al., 2004; Figure 6E
and F) indicated either a symmetric or asymmetric
hand-over-hand mechanism. Thus, the collective SM
data strongly support an asymmetric hand-over-hand
mechanism.
SM Studies of Living Cells
The ultimate goal of all structural, biochemical, and SM
research is to understand the function of biological macromolecules in their natural environments, i.e., in living
cells. The challenge is enormous due to the complexity
of living matter and the relative paucity of appropriate
in vivo methods. SM methods, because of their ability
to follow the behavior of one molecule at a time, seem
to offer an entry into this problem.
Thus far, the prevalent method for observing single
molecules in living cells has been single-dye fluorescence or spFRET. Emission from single fluorophores is
very difficult to detect at the significant background
caused by cellular autofluorescent molecules, such as
flavins, NADH, etc. To minimize such autofluorescence,
fluorophores that are excited by long wavelength (such
as YFP) are needed.
The field is in its infancy, but already some remarkable
successes have been accomplished. Some studies are
summarized in Table 1 (see also Sako and Yanagida
[2003] for an earlier review). Two recent papers from
the laboratory of X.S. Xie demonstrate ingenious ways
to investigate the stochastic expression of mRNA and
protein molecules in individual cells. In one approach
(Cai et al., 2006), bacteria trapped in microcavities exhibit bursts of fluorescent b-galactosidase synthesis:
Pinaud et al., 2004;
for a review, see
Michalet et al. (2005)
Douglass and Vale, 2005
Kural et al., 2005
a few enzyme molecules are produced from an individual mRNA molecule. An equally dramatic result was
obtained by following a fluorescent fusion protein that
localizes to the membrane (Yu et al., 2006). It is clear
that the field has entered an exponential growth phase,
and we can expect a flurry of activity and significant contributions in the not-too-distant future.
Acknowledgments
The authors thank Dr. Steve Block for extensive discussions on the
use of optical tweezers for the study of transcription and for clarifications on the FIONA method. We thank Drs. M. Bennink, T. Strick,
S. Block, A. Yildiz, and P. Selvin for providing original figures and
Dr. A. Thakar for help with figures. J.Z. is supported by NSF Grant
0504239 and start-up funds from the University of Wyoming.
References
Abbondanzieri, E.A., Greenleaf, W.J., Shaevitz, J.W., Landick, R.,
and Block, S.M. (2005). Direct observation of base-pair stepping
by RNA polymerase. Nature 438, 460–465.
Allemand, J.F., Bensimon, D., Lavery, R., and Croquette, V. (1998).
Stretched and overwound DNA forms a Pauling-like structure with
exposed bases. Proc. Natl. Acad. Sci. USA 95, 14152–14157.
Ambrose, W.P., Goodwin, P.M., Martin, J.C., and Keller, R.A. (1994).
Single molecule detection and photochemistry on a surface using
near-field optical excitation. Phys. Rev. Lett. 72, 160–163.
Asbury, C.L., Fehr, A.N., and Block, S.M. (2003). Kinesin moves by an
asymmetric hand-over-hand mechanism. Science 302, 2130–2134.
Ashkin, A. (1997). Optical trapping and manipulation of neutral particles using lasers. Proc. Natl. Acad. Sci. USA 94, 4853–4860.
Axelrod, D. (2001). Total internal reflection fluorescence microscopy
in cell biology. Traffic 2, 764–774.
Bennink, M.L. (2001). Force spectroscopy of single DNA-protein
complexes. An optical tweezers study. PhD thesis, University of
Twente, Enschede, The Netherlands.
Bennink, M.L., Leuba, S.H., Leno, G.H., Zlatanova, J., de Grooth,
B.G., and Greve, J. (2001). Unfolding individual nucleosomes by
Perspective
327
stretching single chromatin fibers with optical tweezers. Nat. Struct.
Biol. 8, 606–610.
Bennink, M.L., Leuba, S.H., and Zlatanova, J. (2005). Analysis of protein/DNA interactions by optical tweezers: application to chromatin
fibers. In Protein-Protein Interactions: A. Molecular Cloning Manual,
E. Golemis and P. Adams, eds. (Cold Spring Harbor, NY: Cold Spring
Harbor Laboratory Press), pp. 415–425.
Best, R.B., and Hummer, G. (2005). Comment on ‘‘Force-clamp
spectroscopy monitors the folding trajectory of a single protein.’’
Science 308, 498.
Binnig, G., Quate, C.F., and Gerber, C. (1986). Atomic force microscope. Phys. Rev. Lett. 56, 930–933.
Brower-Toland, B.D., Smith, C.L., Yeh, R.C., Lis, J.T., Peterson, C.L.,
and Wang, M.D. (2002). Mechanical disruption of individual nucleosomes reveals a reversible multistage release of DNA. Proc. Natl.
Acad. Sci. USA 99, 1960–1965.
Bustamante, C., Macosko, J.C., and Wuite, G.J. (2000a). Grabbing
the cat by the tail: manipulating molecules one by one. Nat. Rev.
Mol. Cell Biol. 1, 130–136.
Bustamante, C., Smith, S.B., Liphardt, J., and Smith, D. (2000b).
Single-molecule studies of DNA mechanics. Curr. Opin. Struct.
Biol. 10, 279–285.
Bustamante, C., Bryant, Z., and Smith, S.B. (2003). Ten years of tension: single-molecule DNA mechanics. Nature 421, 423–427.
Bustamante, C., Chemla, Y.R., Forde, N.R., and Izhaky, D. (2004).
Mechanical processes in biochemistry. Annu. Rev. Biochem. 73,
705–748.
Ha, T. (2004). Structural dynamics and processing of nucleic acids
revealed by single-molecule spectroscopy. Biochemistry 43, 4055–
4063.
Haustein, E., and Schwille, P. (2003). Ultrasensitive investigations of
biological systems by fluorescence correlation spectroscopy.
Methods 29, 153–166.
Haustein, E., and Schwille, P. (2004). Single-molecule spectroscopic
methods. Curr. Opin. Struct. Biol. 14, 531–540.
Herbert, K.M., La Porta, A., Wong, B.J., Mooney, R.A., Neuman, K.C.,
Landick, R., and Block, S.M. (2006). Sequence-resolved detection of
pausing by single RNA polymerase molecules. Cell 125, 1083–1094.
Hua, W., Chung, J., and Gelles, J. (2002). Distinguishing inchworm
and hand-over-hand processive kinesin movement by neck rotation
measurements. Science 295, 844–848.
Journal of Biological Chemistry (1999). Bethesda, MD: American
Society for Biochemistry and Molecular Biology. 274. Minireview series on single-molecules edited by K. van Holde.
Kaseda, K., Higuchi, H., and Hirose, K. (2003). Alternate fast and
slow stepping of a heterodimeric kinesin molecule. Nat. Cell Biol.
5, 1079–1082.
Kelley, A.M., Michalet, X., and Weiss, S. (2001). Chemical physics.
Single-molecule spectroscopy comes of age. Science 292, 1671–
1672.
Khan, S., and Sheetz, M.P. (1997). Force effects on biochemical kinetics. Annu. Rev. Biochem. 66, 785–805.
Knight, A., Mashanov, G., and Molloy, J.E. (2005). Single molecule
measurements and biological motors. Eur. Biophys. J. 35, 89.
Cai, L., Friedman, N., and Xie, X.S. (2006). Stochastic protein expression in individual cells at the single molecule level. Nature 440, 358–
362.
Kural, C., Kim, H., Syed, S., Goshima, G., Gelfand, V.I., and Selvin,
P.R. (2005). Kinesin and dynein move a peroxisome in vivo: a tugof-war or coordinated movement? Science 308, 1469–1472.
Carrion-Vazquez, M., Marszalek, P.E., Oberhauser, A.F., and
Fernandez, J.M. (1999). Atomic force microscopy captures length
phenotypes in single proteins. Proc. Natl. Acad. Sci. USA 96,
11288–11292.
Lakadamyali, M., Rust, M.J., Babcock, H.P., and Zhuang, X. (2003).
Visualizing infection of individual influenza viruses. Proc. Natl. Acad.
Sci. USA 100, 9280–9285.
Cecconi, C., Shank, E.A., Bustamante, C., and Marqusee, S. (2005).
Direct observation of the three-state folding of a single protein molecule. Science 309, 2057–2060.
Charvin, G., Strick, T.R., Bensimon, D., and Croquette, V. (2005).
Tracking topoisomerase activity at the single-molecule level.
Annu. Rev. Biophys. Biomol. Struct. 34, 201–219.
Colquhoun, D., and Hawkes, A.G. (1995). The principles of the stochastic interpretation of ion channel mechanisms. In Single Channel
Recording Second Edition, B. Sakmann and E. Neher, eds. (New
York: Plenum Press), pp. 397–484.
Deniz, A.A., Laurence, T.A., Dahan, M., Chemla, D.S., Schultz, P.G.,
and Weiss, S. (2001). Ratiometric single-molecule studies of freely
diffusing biomolecules. Annu. Rev. Phys. Chem. 52, 233–253.
Douglass, A.D., and Vale, R.D. (2005). Single-molecule microscopy
reveals plasma membrane microdomains created by protein-protein
networks that exclude or trap signaling molecules in T cells. Cell 121,
937–950.
Fernandez, J.M., and Li, H. (2004). Force-clamp spectroscopy monitors the folding trajectory of a single protein. Science 303, 1674–
1678.
Fo¨rster, T. (1959). Transfer mechanisms of electronic excitation.
Discuss. Faraday Soc. 27, 7–17.
Fujiwara, T., Ritchie, K., Murakoshi, H., Jacobson, K., and Kusumi, A.
(2002). Phospholipids undergo hop diffusion in compartmentalized
cell membrane. J. Cell Biol. 157, 1071–1081.
Goulian, M., and Simon, S.M. (2000). Tracking single proteins within
cells. Biophys. J. 79, 2188–2198.
Greenleaf, W.J., Woodside, M.T., Abbondanzieri, E.A., and Block,
S.M. (2005). Passive all-optical force clamp for high-resolution laser
trapping. Phys. Rev. Lett. 95, 208102.
Leckband, D., and Israelachvili, J. (2001). Intermolecular forces in biology. Q. Rev. Biophys. 34, 105–267.
Leuba, S.H., and Zlatanova, J. (2001). Biology at the Single-Molecule
Level (Amsterdam: Pergamon).
Leuba, S.H., Bennink, M.L., and Zlatanova, J. (2003). Single-molecule analysis of chromatin. Methods Enzymol. 376, 73–105.
Lilley, D.M., and Wilson, T.J. (2000). Fluorescence resonance energy
transfer as a structural tool for nucleic acids. Curr. Opin. Chem. Biol.
4, 507–517.
Maier, B., Chen, I., Dubnau, D., and Sheetz, M.P. (2004). DNA transport into Bacillus subtilis requires proton motive force to generate
large molecular forces. Nat. Struct. Mol. Biol. 11, 643–649.
Marszalek, P.E., Oberhauser, A.F., Pang, Y.P., and Fernandez, J.M.
(1998). Polysaccharide elasticity governed by chair-boat transitions
of the glucopyranose ring. Nature 396, 661–664.
Marszalek, P.E., Lu, H., Li, H., Carrion-Vazquez, M., Oberhauser,
A.F., Schulten, K., and Fernandez, J.M. (1999a). Mechanical unfolding intermediates in titin modules. Nature 402, 100–103.
Marszalek, P.E., Pang, Y.P., Li, H., El Yazal, J., Oberhauser, A.F., and
Fernandez, J.M. (1999b). Atomic levers control pyranose ring conformations. Proc. Natl. Acad. Sci. USA 96, 7894–7898.
McKinney, S.A., Tan, E., Wilson, T.J., Nahas, M.K., Declais, A.C.,
Clegg, R.M., Lilley, D.M., and Ha, T. (2004). Single-molecule studies
of DNA and RNA four-way junctions. Biochem. Soc. Trans. 32,
41–45.
Michalet, X., Kapanidis, A.N., Laurence, T., Pinaud, F., Doose, S.,
Pflughoefft, M., and Weiss, S. (2003). The power and prospects of
fluorescence microscopies and spectroscopies. Annu. Rev.
Biophys. Biomol. Struct. 32, 161–182.
Ha, T. (2001a). Single-molecule fluorescence methods for the study
of nucleic acids. Curr. Opin. Struct. Biol. 11, 287–292.
Michalet, X., Pinaud, F.F., Bentolila, L.A., Tsay, J.M., Doose, S., Li,
J.J., Sundaresan, G., Wu, A.M., Gambhir, S.S., and Weiss, S.
(2005). Quantum dots for live cells, in vivo imaging, and diagnostics.
Science 307, 538–544.
Ha, T. (2001b). Single-molecule fluorescence resonance energy
transfer. Methods 25, 78–86.
Murakoshi, H., Iino, R., Kobayashi, T., Fujiwara, T., Ohshima, C.,
Yoshimura, A., and Kusumi, A. (2004). Single-molecule imaging
Molecular Cell
328
analysis of Ras activation in living cells. Proc. Natl. Acad. Sci. USA
101, 7317–7322.
Nakada, C., Ritchie, K., Oba, Y., Nakamura, M., Hotta, Y., Iino, R.,
Kasai, R.S., Yamaguchi, K., Fujiwara, T., and Kusumi, A. (2003).
Accumulation of anchored proteins forms membrane diffusion barriers during neuronal polarization. Nat. Cell Biol. 5, 626–632.
Smith, S.B., Cui, Y., and Bustamante, C. (1996). Overstretching
B-DNA: the elastic response of individual double-stranded and
single-stranded DNA molecules. Science 271, 795–799.
Smith, S.B., Cui, Y., and Bustamante, C. (2003). Optical-trap force
transducer that operates by direct measurement of light momentum.
Methods Enzymol. 361, 134–162.
Neuman, K.C., Abbondanzieri, E.A., Landick, R., Gelles, J., and
Block, S.M. (2003). Ubiquitous transcriptional pausing is independent of RNA polymerase backtracking. Cell 115, 437–447.
Sosnick, T.R. (2004). Comment on ‘‘Force-clamp spectroscopy
monitors the folding trajectory of a single protein.’’ Science 306,
411.
Neuweiler, H., and Sauer, M. (2004). Using photoinduced charge
transfer reactions to study conformational dynamics of biopolymers
at the single-molecule level. Curr. Pharm. Biotechnol. 5, 285–298.
Strick, T.R., Allemand, J.F., Bensimon, D., Bensimon, A., and Croquette, V. (1996). The elasticity of a single supercoiled DNA molecule. Science 271, 1835–1837.
Oberhauser, A.F., Marszalek, P.E., Erickson, H.P., and Fernandez,
J.M. (1998). The molecular elasticity of the extracellular matrix protein tenascin. Nature 393, 181–185.
Strick, T.R., Allemand, J.F., Bensimon, D., and Croquette, V. (1998).
Behavior of supercoiled DNA. Biophys. J. 74, 2016–2028.
Oberhauser, A.F., Marszalek, P.E., Carrion-Vazquez, M., and Fernandez, J.M. (1999). Single protein misfolding events captured by
atomic force microscopy. Nat. Struct. Biol. 6, 1025–1028.
Peterman, E.J., Sosa, H., and Moerner, W.E. (2004). Single-molecule
fluorescence spectroscopy and microscopy of biomolecular motors. Annu. Rev. Phys. Chem. 55, 79–96.
Pinaud, F., King, D., Moore, H.P., and Weiss, S. (2004). Bioactivation
and cell targeting of semiconductor CdSe/ZnS nanocrystals with
phytochelatin-related peptides. J. Am. Chem. Soc. 126, 6115–6123.
Strick, T., Allemand, J., Croquette, V., and Bensimon, D. (2000).
Twisting and stretching single DNA molecules. Prog. Biophys.
Mol. Biol. 74, 115–140.
Stryer, L., and Haugland, R.P. (1967). Energy transfer: a spectroscopic ruler. Proc. Natl. Acad. Sci. USA 58, 719–726.
Tolic´-Nørrelykke, S.F., Engh, A.M., Landick, R., and Gelles, J. (2004).
Diversity in the rates of transcript elongation by single RNA polymerase molecules. J. Biol. Chem. 279, 3292–3299.
Tsien, R.Y. (1998). The green fluorescent protein. Annu. Rev. Biochem. 67, 509–544.
Revyakin, A., Ebright, R.H., and Strick, T.R. (2005). Single-molecule
DNA nanomanipulation: improved resolution through use of shorter
DNA fragments. Nat. Methods 2, 127–138.
Ueda, M., Sako, Y., Tanaka, T., Devreotes, P., and Yanagida, T.
(2001). Single-molecule analysis of chemotactic signaling in Dictyostelium cells. Science 294, 864–867.
Rief, M., Gautel, M., Oesterhelt, F., Fernandez, J.M., and Gaub, H.E.
(1997). Reversible unfolding of individual titin immunoglobulin domains by AFM. Science 276, 1109–1112.
van Holde, K.E., Johnson, W.C., and Ho, P.S. (2006). Principles of
Physical Biochemistry (Upper Saddle River, NJ: Pearson/Prentice
Hall).
Sabanayagam, C.R., Eid, J.S., and Meller, A. (2005). Long time scale
blinking kinetics of cyanine fluorophores conjugated to DNA and its
effect on Forster resonance energy transfer. J. Chem. Phys. 123,
224798.
Sako, Y., and Yanagida, T. (2003). Single-molecule visualization in
cell biology. Nat. Rev. Mol. Cell Biol. (Suppl), SS1–SS5.
Sako, Y., Hibino, K., Miyauchi, T., Miyamoto, Y., Ueda, M., and Yanagida, T. (2000a). Single-molecule imaging of signaling molecules in
living cells. Single Mol. 1, 151–155.
Sako, Y., Minoghchi, S., and Yanagida, T. (2000b). Single-molecule
imaging of EGFR signalling on the surface of living cells. Nat. Cell
Biol. 2, 168–172.
Sarkar, A., Robertson, R.B., and Fernandez, J.M. (2004). Simultaneous atomic force microscope and fluorescence measurements
of protein unfolding using a calibrated evanescent wave. Proc.
Natl. Acad. Sci. USA 101, 12882–12886.
Schafer, D.A., Gelles, J., Sheetz, M.P., and Landick, R. (1991). Transcription by single molecules of RNA polymerase observed by light
microscopy. Nature 352, 444–448.
Schmoranzer, J., Goulian, M., Axelrod, D., and Simon, S.M. (2000).
Imaging constitutive exocytosis with total internal reflection fluorescence microscopy. J. Cell Biol. 149, 23–32.
Schutz, G.J., Kada, G., Pastushenko, V.P., and Schindler, H. (2000).
Properties of lipid microdomains in a muscle cell membrane visualized by single molecule microscopy. EMBO J. 19, 892–901.
Science (1999). Washington, DC: Science/AAAS. 283. March issue
devoted to reviews on single-molecule approaches.
Seisenberger, G., Ried, M.U., Endress, T., Buning, H., Hallek, M., and
Brauchle, C. (2001). Real-time single-molecule imaging of the infection pathway of an adeno-associated virus. Science 294, 1929–1932.
Selvin, P.R. (2000). The renaissance of fluorescence resonance energy transfer. Nat. Struct. Biol. 7, 730–734.
Shaevitz, J.W., Abbondanzieri, E.A., Landick, R., and Block, S.M.
(2003). Backtracking by single RNA polymerase molecules observed
at near-base-pair resolution. Nature 426, 684–687.
Smith, S.B., Finzi, L., and Bustamante, C. (1992). Direct mechanical
measurements of the elasticity of single DNA molecules by using
magnetic beads. Science 258, 1122–1126.
Visscher, K., and Block, S.M. (1998). Versatile optical traps with
feedback control. Methods Enzymol. 298, 460–489.
Watanabe, N., and Mitchison, T.J. (2002). Single-molecule speckle
analysis of actin filament turnover in lamellipodia. Science 295,
1083–1086.
Weiss, S. (1999). Fluorescence spectroscopy of single biomolecules. Science 283, 1676–1683.
Weiss, S. (2000). Measuring conformational dynamics of biomolecules by single molecule fluorescence spectroscopy. Nat. Struct.
Biol. 7, 724–729.
Xie, X.S., and Dunn, R.C. (1994). Probing single molecule dynamics.
Science 265, 361–364.
Yang, W., Gelles, J., and Musser, S.M. (2004). Imaging of singlemolecule translocation through nuclear pore complexes. Proc.
Natl. Acad. Sci. USA 101, 12887–12892.
Yildiz, A., and Selvin, P.R. (2005a). Fluorescence imaging with one
nanometer accuracy: application to molecular motors. Acc. Chem.
Res. 38, 574–582.
Yildiz, A., and Selvin, P.R. (2005b). Kinesin: walking, crawling or sliding along? Trends Cell Biol. 15, 112–120.
Yildiz, A., Forkey, J.N., McKinney, S.A., Ha, T., Goldman, Y.E., and
Selvin, P.R. (2003). Myosin V walks hand-over-hand: single fluorophore imaging with 1.5-nm localization. Science 300, 2061–2065.
Yildiz, A., Tomishige, M., Vale, R.D., and Selvin, P.R. (2004). Kinesin
walks hand-over-hand. Science 303, 676–678.
Yin, H., Landick, R., and Gelles, J. (1994). Tethered particle motion
method for studying transcript elongation by a single RNA polymerase molecule. Biophys. J. 67, 2468–2478.
Yu, J., Xiao, J., Ren, X., Lao, K., and Xie, X.S. (2006). Probing gene
expression in live cells, one protein molecule at a time. Science
311, 1600–1603.
Zheng, H., Tomschik, M., Zlatanova, J., and Leuba, S.H. (2005).
Evanescent field fluorescence microscopy for analysis of protein/
DNA interactions at the single-molecule level. In Protein-Protein Interactions: A. Molecular Cloning Manual, E. Golemis and P. Adams,
eds. (Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press),
pp. 429–444.
Zlatanova, J. (2003). Forcing chromatin. J. Biol. Chem. 278, 23213–
23216.
Perspective
329
Zlatanova, J., and Leuba, S.H. (2002). Stretching and imaging single
DNA molecules and chromatin. J. Muscle Res. Cell Motil. 23, 377–
395.
Zlatanova, J., and Leuba, S.H. (2003a). Chromatin fibers, one-at-atime. J. Mol. Biol. 331, 1–19.
Zlatanova, J., and Leuba, S.H. (2003b). Magnetic tweezers: a sensitive tool to study DNA and chromatin at the single-molecule level.
Biochem. Cell Biol. 81, 151–159.
Zlatanova, J., and Leuba, S.H. (2004). Chromatin structure and
dynamics: lessons from single molecule approaches. In Chromatin
Structure and Dynamics: State-of-the-Art., J. Zlatanova and S.H.
Leuba, eds. (Amsterdam: Elsevier), pp. 309–342.
Zlatanova, J., Lindsay, S.M., and Leuba, S.H. (2000). Single molecule
force spectroscopy in biology using the atomic force microscope.
Prog. Biophys. Mol. Biol. 74, 37–61.
Zlatanova, J., McAllister, W.T., Borukhov, S., and Leuba, S.H. (2006).
Single-molecule approaches reveal the idiosyncrasies of RNA polymerases. Structure 14, 953–966.