ChapterPDF Available

An Introduction to Magnetic Tweezers

Authors:

Abstract and Figures

Magnetic tweezers are a single-molecule force and torque spectroscopy technique that enable the mechanical interrogation in vitro of biomolecules, such as nucleic acids and proteins. They use a magnetic field originating from either permanent magnets or electromagnets to attract a magnetic particle, thus stretching the tethering biomolecule. They nicely complement other force spectroscopy techniques such as optical tweezers and atomic force microscopy (AFM) as they operate as a very stable force clamp, enabling long-duration experiments over a very broad range of forces spanning from 10 fN to 1 nN, with 1–10 milliseconds time and sub-nanometer spatial resolution. Their simplicity, robustness, and versatility have made magnetic tweezers a key technique within the field of single-molecule biophysics, being broadly applied to study the mechanical properties of, e.g., nucleic acids, genome processing molecular motors, protein folding, and nucleoprotein filaments. Furthermore, magnetic tweezers allow for high-throughput single-molecule measurements by tracking hundreds of biomolecules simultaneously both in real-time and at high spatiotemporal resolution. Magnetic tweezers naturally combine with surface-based fluorescence spectroscopy techniques, such as total internal reflection fluorescence microscopy, enabling correlative fluorescence and force/torque spectroscopy on biomolecules. This chapter presents an introduction to magnetic tweezers including a description of the hardware, the theory behind force calibration, its spatiotemporal resolution, combining it with other techniques, and a (non-exhaustive) overview of biological applications.
Content may be subject to copyright.
Chapter 18
An Introduction to Magnetic Tweezers
David Dulin
Abstract
Magnetic tweezers are a single-molecule force and torque spectroscopy technique that enable the mechani-
cal interrogation in vitro of biomolecules, such as nucleic acids and proteins. They use a magnetic field
originating from either permanent magnets or electromagnets to attract a magnetic particle, thus stretching
the tethering biomolecule. They nicely complement other force spectroscopy techniques such as optical
tweezers and atomic force microscopy (AFM) as they operate as a very stable force clamp, enabling
long-duration experiments over a very broad range of forces spanning from 10 fN to 1 nN, with
110 milliseconds time and sub-nanometer spatial resolution. Their simplicity, robustness, and versatility
have made magnetic tweezers a key technique within the field of single-molecule biophysics, being broadly
applied to study the mechanical properties of, e.g., nucleic acids, genome processing molecular motors,
protein folding, and nucleoprotein filaments. Furthermore, magnetic tweezers allow for high-throughput
single-molecule measurements by tracking hundreds of biomolecules simultaneously both in real-time and
at high spatiotemporal resolution. Magnetic tweezers naturally combine with surface-based fluorescence
spectroscopy techniques, such as total internal reflection fluorescence microscopy, enabling correlative
fluorescence and force/torque spectroscopy on biomolecules. This chapter presents an introduction to
magnetic tweezers including a description of the hardware, the theory behind force calibration, its
spatiotemporal resolution, combining it with other techniques, and a (non-exhaustive) overview of
biological applications.
Key words Single-molecule biophysics, Magnetic tweezers, Force spectroscopy, Protein-nucleic acids
interactions, Torque spectroscopy
1 Brief History and Application of Magnetic Tweezers
Magnetic tweezers use the magnetic field generated by permanent
or electromagnets to apply force and/or rotate magnetic particles
attached to a biological material, hence inducing a mechanical
stress. The first biophysics assay using a magnetic actuator in a
biological context was reported by Crick and Hughes in 1950
[1], where magnetic particles placed in the cytoplasm of a cell
were displaced to interrogate its viscoelastic properties. Magnetic
tweezers have two main areas of application in biophysics: cellular
mechanics and single-molecule biophysics. The force applied in
Iddo Heller et al. (eds.), Single Molecule Analysis: Methods and Protocols, Methods in Molecular Biology, vol. 2694,
https://doi.org/10.1007/978-1-0716-3377-9_18,
© The Author(s) 2024
375
cellular mechanics is usually relatively large ( 1 nN) [2, 3] com-
pared to the single-molecule world (100 pN) [4], and therefore
their respective instrument designs differ significantly. Here, we
solely focus on magnetic tweezers assays for single-molecule
biophysics.
376 David Dulin
In the 1990s, the Bustamante lab and the Croquette and
Bensimon lab pioneered the modern version of single-mole-
cule magnetic tweezers capable of applying a constant force (even
below 1 pN) and torque, enabling the interrogation of DNA
mechanical properties at the single-molecule level [5,6]. Nowadays,
magnetic tweezers are found in many labs around the world, and
single-molecule studies have been performed on various protein-
nucleic acids systems [7], ranging from helicases [8] to DNA poly-
merases [79], topoisomerases and gyrases [1013], cellular and
viral RNA polymerases [14], nucleoprotein filaments [1518], and
the mechanical stability of protein folding and protein-ligand inter-
actions [1923]. The simplicity and the robustness of the technique
make it a powerful single-molecule force spectroscopy assay that
has become more and more popular in the academic community.
2 Description of a Magnetic Tweezers Apparatus
Magnetic tweezers for single-molecule studies are composed of a
collimated light source located above a magnetic field source (e.g.,
permanent magnets) that is mounted on top of a flow cell in which
super-paramagnetic beads (simply coined magnetic beads from
now on) are tethered to the flow cell coverslip surface by a biomol-
ecule (Fig. 1a). The magnetic beads are imaged using an inverted
microscope onto a camera, which enables the tracking of their
three-dimension position as a function of time. The latest comple-
mentary metal-oxide-semiconductor (CMOS) cameras enable both
high-throughput (Fig. 1b)[2427] and high-speed measurements
[2830]. The vertical and angular position of the magnets is
adjusted using linear motors to vary the force (Fig. 1c) and the
torque (Fig. 1d), respectively, applied to the biomolecule. The
inverted microscope body may be either custom build or bought
commercially. Given its simple design, the custom body presents
only a mild difficulty to produce and is mechanically more stable.
2.1 Magnet
Configuration
Different configurations have been used in magnetic tweezers
experiments to modulate how force and torque are applied
[15]. The most standard configuration uses a pair of cubic perma-
nent magnets, being either vertically (Fig. 1a, c, and d) or horizon-
tally aligned [31]. While the gap between the magnets gives access
to the light source, it also modulates the applied force: a smaller
gap results in a larger maximum force but reduces the surface
area that experiences a homogenous force field [32, 33]. The
two-permanent magnet cubes configuration strongly clamps the
magnetic bead in rotation, which precludes specific applications,
e.g., directly measuring the torsional stiffness of a soft biomolecule
such as DNA. Alternative magnet geometries have therefore been
developed using either a single cylindrical magnet [34] or addition-
ally having a small side magnet attached to it to angularly trap the
magnetic bead mildly [35, 36]. These methods have been reviewed
in detail in Ref. [15].
An Introduction to Magnetic Tweezers 377
Fig. 1 Magnetic tweezers for single-molecule applications. (a) Schematic of a magnetic tweezers instrument
(Adapted from Ref. [61]). (b) Field of view of a high-throughput magnetic tweezers assay, where ~500
magnetic beads of 2.8 μm diameter can be followed simultaneously (50× magnification, 120 nm pixel size)
(Adapted from Ref. [49]). (c, d) Force and torque spectroscopy, respectively, of a single nucleic acid molecule
using magnetic tweezers. The nucleic acid molecule is attached to the magnetic bead via a biotin-streptavidin
bond, and to the surface via digoxigenin-antidigoxigenin attachment. Biotin and digoxigenin molecules are
inserted nonspecifically during the synthesis of the nucleic acid handles that are subsequently ligated to the
main nucleic acid strand (Adapted from Ref. [51])
378 David Dulin
2.2 Illumination Different light sources have been used to illuminate the sample. A
source is chosen to be spatially and temporally coherent, which
generates many diffraction rings with a good contrast to enable
an excellent tracking resolution. Light-emitting diodes (LEDs) are
a simple solution that provide a good temporal coherence (~10 nm
spectral dispersion, full width at half maximum), are easily colli-
mated using a high numerical aperture aspherical lens, and provide
enough light intensity to image the beads in standard image acqui-
sition frequency (50100 Hz) (Fig. 1a). To achieve high-speed
image acquisition ( kHz), a high photon flux through the sample
is required. Unfortunately, LEDs can hardly satisfy this require-
ment. With a noncoherent light source, a high flux may be achieved
using a fiber-coupled arc lamp in combination with a spectral filter
[30, 37]. Coherent sources such as laser diodes and super-
luminescent diodes also enable an efficient collection and collima-
tion of the output light onto the sample [38]. This enables short
camera shutter times and therefore high image acquisition rates.
However, these coherent sources have the shortcoming of creating
spurious speckle patterns in the field of view. Dark field illumina-
tion, i.e., by blocking the zero-order light pathway, has recently
demonstrated the best to date resolution by reducing the back-
ground noise significantly [39].
2.3 Bead Position
Tracking Algorithm
To follow biomolecular reactions with magnetic tweezers, one must
precisely track the magnetic bead’s position in three dimensions. To
this end, different algorithms have been developed, all using the
diffraction pattern originating from the out-of-focus micron-sized
beads (Fig. 2ac). A region of interest around the bead is defined a
priori to indicate which area of the camera image contains single,
insolated beads (Fig. 2a, b). A lookup table is acquired before the
start of the experiment by capturing a diffraction pattern of the bead
at different objective positions along the z-axis (Fig. 2a)
[40, 41]. The objective is displaced using a high-resolution piezo
stage in steps of ~50100 nm (Fig. 1a). To determine the bead’s
position in the (x,y)-plane, a rough estimate is obtained from a center
of mass, followed by a cross-correlation algorithm (Fig. 2d). More
sophisticated versions of this tracking algorithm have been devel-
oped, such as the quadrant interpolation method [25, 42]. To deter-
mine the axial position (z-axis), the diffraction pattern of the
magnetic bead at any given frame is compared to the lookup table
using a squared error metric. Sub-plane resolution is then obtained
by a polynomial fit of the resulting error curve (Fig. 2a). Because of
the simplicity of these tracking algorithms, hundreds of beads may be
followed in parallel and in real-time using modern GPUs to perform
the calculation [25, 30].
An Introduction to Magnetic Tweezers 379
Fig. 2 Bead localization in a magnetic tweezers experiment. (a) Left, diffraction
pattern of a 3 μm diameter reference bead attached to the surface of a coverslip
at different distances from the bead to the microscope objective’s focal plane
along the z-axis. Right, a lookup table built up from radial intensity profiles
across the center of the images of the dif fraction pattern taken at different focal
plane positions spaced by 50 nm intervals. (b, c) Diffraction pattern and intensity
profile along the x-axis (black line). (d) Autocorrelation function (ACF) between
two profiles as in (b) separated by 10 pixels (gray). The maximum of ACF is
indicated by the arrow
380 David Dulin
2.4 Temperature
Control
Enzymatic reactions are sensitive to temperature fluctuations and
follow the Arrhenius law:
kTðÞ=Ae -Ea=kBT ,ð1Þ
where k is the forward reaction rate constant, A is a pre-exponential
factor, E
a
is the activation energy, k
B
the Boltzmann constant, and
T the temperature [43]. It is therefore of great importance to
precisely control the temperature in the flow chamber. Several
articles have been reported on establishing a temperature control
on the flow chambers [4446]. Simulation and data have clearly
demonstrated that the main heat sink is the oil immersion objective
[45], which directly contacts with the glass coverslip area where the
reaction occurs. In conclusion, controlling the temperature at the
objective enables a precise control of the temperature of the reac-
tion (±0.1 °C), which can be achieved using a simple device com-
mercially available from Thorlabs [46].
2.5 Surface
Functionalization and
Nucleic Acid Construct
Fabrication
Magnetic tweezers are a surface-based technique (Fig. 1a); there-
fore, the flow chamber’s glass surface should be treated with care to
prevent nonspecific attachment to the surface of either the mag-
netic beads or the biomolecules of interest. Different types of
surface functionalization have been developed, such as polyethyl-
ene glycol (PEG) [47, 48], nitrocellulose [49], and lipid bilayer
[50]. The last two are being described in detail in Chapter 21. The
type of attachment is of great importance and therefore defines the
methodology to generate the tether. The standard method of
tethering the magnetic beads relies on fabricating nucleic acids
containing both a biotin handle on one end, to attach the strepta-
vidin coated bead, and a digoxigenin (dig) handle on the other end
to attach the nucleic acid to anti-digoxigenin (anti-dig) antibodies
adsorbed to the flow chamber’s glass surface [51]. Such labels are
introduced when generating the nucleic acid by adding dig- or
biotin-labeled UTP to the nucleotide sets. While biotin-
streptavidin forms a very stable bond, the dig-anti-dig bond is
much weaker and not suitable for high-force or long experiments,
even when using glutaraldehyde to cross-link proteins to the nitro-
cellulose surface [52]. For such experiments, covalent attachment is
preferred to replace the dig-anti-dig bond using either a PEG
functionalized surface with covalent chemistry to attach the bio-
molecule to the surface [47, 48] or a direct attachment [ 20],
providing a tether with a much longer tether surface attachment
lifetime .
DNA and RNA construct fabrication rely either on specific
ligation of double-stranded ends (Fig. 3a, b), annealing single-
stranded nucleic acids, or a combination of both. Very detailed
protocols can be found in several method articles [5156], and
this topic will therefore not be further discussed here.
An Introduction to Magnetic Tweezers 381
Fig. 3 DNA construct fabrication for single-molecule force spectroscopy experiments. (a) Steps in synthesizing
double-stranded DNA constructs. A plasmid is digested to generate a stem. Handles labeled with either biotins
(BIO) or digoxigenins (DIG) are generated by PCR using λ phage DNA as a template and by adding either
bio-dUTP or dig-dUTP in the reaction solution. The handles are purified, digested, and ligated to the stem. (b)
Similar approach as in (a) to fabricate a DNA hairpin. The different segments are produced by PCR digestion
and ligated together to shape as a hairpin
3 Physical Principles
3.1 Force and Torque
Origin
Magnetic tweezers can apply forces between a femto-Newton
(fN) and a nano-Newton (nN) [4], which depends on the magnetic
bead size (i.e., the total amount of magnetic content) and the
magnet configuration. In the configuration described in Fig. 1a,
reducing the gap between the two magnets increases the force.
Because of the very large force range accessible, magnetic tweezers
have been applied to investigate very different biomolecular sys-
tems. The force experienced by a magnetic particle in a magnetic
field is described by:
F
mag = 1
2
m
sat B
,ð2Þ
where F mag is the magnetic force, m
sat is the saturated magne-
tization of the particle, and B
is the magnetic field [32]. Interest-
ingly, the magnetic force is directly proportional to the gradient of
the magnetic field, not its magnitude.
One of the key aspects of magnetic tweezers is their ability to
apply torque to the tether [6, 57], and torque spectroscopy was
performed on many different complexes, e.g., double-stranded
nucleic acids and nucleoprotein filaments [15]. The magnetic beads
are made of super-paramagnetic nanoparticles embedded in a latex
matrix, and therefore their magnetization should align with the
magnetic field. However, an asymmetry in the nanoparticle spatial
organization induces an anisotropy in m
sat, with a minor compo-
nent m
0 not aligned with B
[58]. This induces a torque Γ
on the
bead, which is derived from:
382 David Dulin
Γ
= m
0 × B
:ð3Þ
The torque response of the biomolecule is negligible in respect
to Γ
. Hence, rotating the magnets induces a rotation of the mag-
netic bead, thereby transferring torque to the coilable tether. An
example of coilable tether is a fully double-stranded nucleic acid
molecule with multiple attachment points at both ends (Fig. 1a).
To measure the torque response of the biomolecule, and therefore
its torsional stiffness and related mechanical properties, different
magnet configurations have been developed to reduce the magni-
tude of Γ
, such as the magnetic torque tweezers [35].
3.2 Force Calibration The force F
mag
may be calibrated from Eq. (1), by using m
sat
(from
the factory specifications of the magnetic beads) and spatially char-
acterizing the magnetic field generated by the magnets with a Hall
probe [32]. However, this method is not the preferred one, as it
relies on parameters measured externally for a given batch of beads.
Therefore, in situ force calibration methods that rely on parameters
directly measured in the magnetic tweezers assay are preferred
[59]. To this end, the theory relating the force as a function of
the tethered magnetic bead lateral fluctuations is derived below.
The force may therefore be extracted from measuring such
fluctuations.
The position of a tethered magnetic bead experiencing a force
F
mag
is best described as an inverted pendulum (Fig. 4a, b)[6, 60]
with two representative cases of pendulum lengths coined short
(Fig. 4a) and long pendulum (Fig. 4b), respectively. In the former
case, the fluctuation in position along the x-axis is pinned by the
magnetic field B
(Fig. 4a), and the length of the pendulum is
therefore the length of the tether L
ext
. In the latter case, the
fluctuation in position of the bead along the y-axis is not con-
strained by the magnetic field (Fig. 4a), and the length of the
pendulum is therefore L
ext
+ R, where R is the magnetic bead
radius.
Considering the short pendulum case, the small displacement
δx from the equilibrium position caused by the collisions with the
water molecules (Fig. 4c), i.e., the Brownian motion, induces a
restoring force that can be described as:
Frestoring =kxδx,ð4Þ
h i
An Introduction to Magnetic Tweezers 383
Fig. 4 Force calibration in a magnetic tweezers instrument. (a, b) Schematic of the tethered magnetic bead
position fluctuations along either (a) the x-axis (short pendulum) or (b) the y-axis (long pendulum). (c)
Schematic of the forces exerted on the magnetic bead in the short pendulum configuration. (d) Position of
a magnetic bead along the x-axis against the y-position (left) and time (right). Data taken at three different
forces. (e) Force calibration for M270 (blue, 2.8 μm diameter) and MyOne (red, 1 μm diameter) magnetic
beads as a function of the distance of the magnets to the flow chamber (Adapted from Ref. [61])
where k
x
is the trap stiffness along the x-axis. Therefore,
Frestoring =Fmag sin θðÞ=Fmag δx=Lext,ð5Þ
and
kx =Fmag =Lext,ð6Þ
where θ is the angle spanned by the tether when the bead is at its
current and equilibrium positions, and L
ext
is the length of tether.
At small θ, the potential energy landscape U
x
is quadratic [60], i.e.,
Ux = 1=2ðÞkxδx2 ,ð7Þ
and we can therefore apply the equipartition theorem
( U
x
= (1/2) k
B
T) on Eq. (7), and we obtain:
ÞÞ
ÞÞ
384 David Dulin
hδx2i=kBT=kx =kBTLext=Fmag ,ð8Þ
and, by extension, for the long pendulum case:
δy2 =kBT=ky =kBTLext þ RðÞ=Fmag:ð9Þ
Equation (8) and (9) directly link the applied force with the
fluctuations in the lateral position of the bead and the tether length
(Fig. 4d). Both are parameters that one can easily retrieve from
experiments to enable a direct force calibration as a function of the
distance of the magnets from the magnetic bead (Fig. 4e)[61, 62].
To provide an accurate force calibration, the lateral fluctuation
of the bead must be measured accurately to not overestimate the
force (Eqs. 8 and 9). To this end, one should make sure the image
acquisition does not overly integrate (meaning average) the mag-
netic bead position fluctuation [61]. From the equation of motion
of the magnetic bead experiencing F
mag
, we are able to extract the
characteristic time scale of the bead:
tc,x =γ=kx =6πηRLext=Fmag and tc,y =γ=ky
=6πηRL
ext þ RðÞ=Fmag,ð10Þ
where γ is the drag coefficient, η the viscosity of the solution
(typically water, i.e., ~10
-3
Pa.s), and R the radius of the magnetic
bead and defines the time during which the bead has explored the
trap. For L
ext
R, the drag coefficient must be corrected to include
the effect of the surface, as described by the Faxe
´n law [63]:
γFaxen =6πηR= 1-9=8 R= RþLext
ðÞðÞþ1=2 R= RþLext
ðÞðÞ
3-57=100 R= RþLext
ðð4
þ1=5 R= RþLext
ðÞðÞ
5þ7=200 R= RþLext
ðÞðÞ
11 -1=25 R= RþLext
ðð12 ,
ð11Þ
hδx
2
i averages away toward a measured value hδx
2
i
meas
as a
function of the camera shutter time τ
sh
and t
c, x
as
δx2
meas = 2kBT=πkx
ðÞarctan 4πtc,x=τsh :ð12Þ
To minimize the error in the force due to camera image blur-
ring, we must minimize the difference between hδx
2
i and hδx
2
i
meas
.
For example, to measure F
mag
with a 10% error due to camera
image blurring, τ
sh
must be at least four times smaller than t
c, x
[61]. How feasible is this in practice? Most large chip CMOS
cameras acquire images with a frequency f
ac
~10 -100 Hz, while
the characteristic time for a L
ext
= 1 μm, R = 1.4 μm and
F
mag
= 10 pN (typical experimental conditions) is t
c, x
~0.03 s,
i.e., similar to τ
sh
for zero-dead time image acquisition ( f
ac
~1/τ
sh
).
In such case, one may use longer DNA tethers to increase t
c, x
in
respect of τ
sh
and extract a calibration table for F
mag
as a function of
the magnets distance to the magnetic bead [59]. However, this
only works for magnetic beads with a small dispersion in magnetic
content, hence in force, such as the Dynabeads M-270
(R = 1.4 μm) and MyOne (R = 0.5 μm) magnetic beads from
Invitrogen. For shorter tethers or higher forces, one may use a very
fast camera, i.e., f
ac
in the kilohertz range, or use a nonzero dead
time acquisition, i.e., τ
sh
1/f
ac
[59, 61]. The former is not
available for all camera models and only when using a small field
of view [2830].The latter is easily programmable in most cameras
without compromising the field of view [61]. Another possibility is
to correct hδx
2
i
meas
for the camera image blurring, either in the
frequency or time domain [59, 64]. This works well for τ
sh
/2 < t
c,
x
< τ
sh
/4 [60], and packages in MATLAB and Python are available
to perform such calibrations [59, 65]. These strategies however fail
to perform accurate force calibration for very short tethers, as the
rotation of the bead induced by the magnetic field pinning by the
magnetic bead must be accounted for [66, 67]. Similar strategies to
calibrate the force may also be applied to acoustic force spectros-
copy (AFS) [68], as the tethered bead is described by a similar
model (i.e., the inverted pendulum).
An Introduction to Magnetic Tweezers 385
3.3 Estimating the
Spatiotemporal
Resolution of Magnetic
Tweezers
The main parameters measured in magnetic tweezers experiments
are the change in the tether’s extension L
ext
due to either a mechan-
ical response of the tether or an enzymatic activity modifying the
tether length. It is therefore essential to determine the noise ampli-
tude along the z-axis. The spatiotemporal resolution in a magnetic
tweezers assay depends on the tracking and thermal noise as
follows:
δztot
hi
= δztr 2 þ δzth 2:ð13Þ
3.3.1 Tracking
Resolution and Stability
The tracking resolution is defined by the hardware (microscope
objective magnification, numerical aperture, pixel size and light
intensity) and the algorithm used. To experimentally evaluate δz
tr
,
the Allan deviation (AD) is particularly useful [28, 64, 69]
(Fig. 5a). The AD of a particle position along the, e.g., z-axis, is
defined as follows:
σAD τðÞ= 1
2 zτ,jþ1-zτ,j
2 with zτ,j = 1
τ
τ jþ0:5ðÞ
τ j-0:5ðÞ
ztðÞdt, ð14Þ
where τ defines both the time between consecutive samples and the
time over which the sample is averaged. Simply put, the AD is
one-half the average difference in position between consecutive
intervals of duration τ, averaged over all intervals of duration τ.
For a bead stuck to the surface, we observe two regimes (Fig. 5a):
AD initially decreases as 1= τ
p , indicating how the frame-to-frame
uncorrelated noise averages out, and the AD subsequently reaches a
lower bound and rises again due to long timescale drift dominating
the noise (e.g., mechanical drift, tracking algorithm bias). To
improve the stability during the measurement, the mechanical
drift is corrected by subtracting the position of a reference bead
fixed to the flow chamber surface from the position of the magnetic
bead (Figs. and )[ ]. The resolution of the bead position
tracking may be further improved by setting an autofocus locked
onto the position of a reference bead and adjusting the objective’s
285a1a
386 David Dulin
Fig. 5 Spatiotemporal resolution of a magnetic tweezers instrument. (a) Allan deviation (AD) of the z-axis
position of a 3 μm diameter surface-attached polystyrene reference bead (blue), subtracted to another
reference bead (RS, orange), and using autofocus (AF, green). The data were acquired using a 100× objective
magnification and at 58 Hz acquisition frequency. (b) Raw (gray) and 1 Hz low -pass filtered (dark gray) trace
acquired while using the autofocus and drift corrected by subtracting the z-position of another reference bead
(green in (a)) (Adapted from Ref. [49]). (c, d) Height of a (c) reference bead and (d) DNA-tethered bead while
using the piezo stage to move the sample by the increments indicated on top of the panel (in nm) (Adapted
from Ref. [69]). (e, f) Magnetic tweezers assay to monitor single-nucleotide steps of Upf1 helicase when
unwinding a DNA hairpin (Adapted from Ref. [39])
ÞÞ
focal plane’s position using a high-resolution piezo stage and
increase the τ at which AD rises again by decreasing the negative
impact of the tracking algorithm bias [49] (Fig. 5a, b). For a
magnetic tweezers instrument with 100× magnification, 1.25
numerical aperture microscope objective, and 60 nm pixel size in
the image plane, tracking resolutions for a single image of δz
tr
~1 nm
and δz
tr
~0.3 nm are achievable for 1 μm and 3 μm diameter beads,
respectively (using the quadrant interpolation algorithm) [69].
An Introduction to Magnetic Tweezers 387
The tracking resolution may be improved by acquiring data at
high f
ac
and subsequently averaging out the tracking noise by
integrating the bead position over N frames. This results in a
reduction of the tracking noise by a factor of N
p , enabling the
observation of steps as small as 0.3 nm for a reference bead with the
standard magnetic tweezers configuration (Fig. 5c)[2830, 69]
and for a tethered magnetic bead (Fig. 5d)[69]. Recent develop-
ments in magnetic tweezers instrumentation, specifically in the
illumination and imaging path, have enabled the first observation
of single-nucleotide translocation steps by a helicase unwinding a
DNA hairpin (Fig. 5e, f)[39].
3.3.2 Thermal Noise The thermal noise depends on the tether stiffness, k
z
, as follows:
δz2
th =kBT=kz,ð15Þ
where
kz =
Fmag Lext
ðÞ
Lext
= kBT
2LPLC 2þ 1-Lext
LC ð16Þ
with L
P
and L
C
being the persistence and the contour length of the
tether, respectively, assuming the response of the tether to F
mag
is
well-described by the inextensible Worm-like chain model
[70]. Similar to hδx
2
i
meas
, the thermal noise is integrated by the
camera during the image acquisition. Hence,
δz2
th meas = 2kBT=πkz
ðÞ arctan 4πtc,z=τsh ð17Þ
with t
c, z
= γ/k
z
. For L
ext
R, γ must be corrected to account for
the coverslip surface effect using Brenner’s approximation [63]:
γBrenner =6πηR= 1-9=8 R= Rþ Lext
ðÞðÞþ 3=8 R= Rþ Lext
ðð3
-1=4 R= Rþ Lext
ðÞðÞ
4 ,
ð18Þ
Ideally, the resolution of the magnetic tweezers assay is limited
by the thermal noise, which can be estimated using Eqs. 16 and 17.
An accurate simulation of the overall measurement noise for a
tethered magnetic bead in a magnetic tweezers assay has been
described by Burnham and colleagues [71], which is useful to
estimate the spatiotemporal resolution for a given magnetic twee-
zers experiment.
388 David Dulin
Fig. 6 DNA supercoiling experiments using magnetic tweezers. Dynamic
rotation-extension experiment on a 21 kbp long DNA tether at either 0.3 pN
(gray) or 4 pN (black). At low force, both negative and positive supercoils induce
plectonemes. At high force, positive supercoils induce plectonemes, while
negative supercoils unwind the DNA tether. The arrow indicates the buckling
transition at high force
3.4 Using Torque
Spectroscopy in
Magnetic Tweezers
One of the key aspects of magnetic tweezers is their ability to
control the torque applied to a coilable biomolecule (Fig. 1d,
Fig. 6)[6, 57]. This has been (and still is) used to investigate the
mechanical response to torque of double-stranded nucleic acids
[15, 72]. To enable torque spectroscopy, the nucleic acid must be
topologically constrained, i.e., without free rotation point, such as a
fully double-stranded DNA with multiple attachment points at
both ends (Fig. 1d). The twist (Tw) and the writhe (Wr) define
the supercoiled state of the molecule. The former is the number of
times the molecule turns around itself, such as for the DNA double
helix, and the latter is defined by the number of times the molecule
winds over itself. The helical pitch for a relaxed DNA molecule is
10.5 bp/turn and, therefore, the total twist in a relaxed DNA
molecule (Tw
0
) is the number of base pairs divided by the helical
pitch. The linking number (Lk), which is the sum of twist and
writhe, is a topological invariant for a torsionally constrained mole-
cule, meaning
Lk =Tw þ Wr =constant:ð19Þ
For a torsionally relaxed DNA molecule, Wr = 0, so Lk
0
= Tw
0
.
A molecule is said to be supercoiled when Lk Lk
0
.
In addition, the supercoil density σ is a useful description of the
torsional state of a molecule:
An Introduction to Magnetic Tweezers 389
σ = Lk -Lk0
ðÞ=Lk0:ð20Þ
L
ext
as a function of σ is often used to represent magnetic
tweezers experiments investigating the response of double-
stranded nucleic acids to torsional stress. This provides an easy
way to compare the torsional properties of DNA tethers of different
lengths.
Upon addition of positive turns to a torsionally relaxed mole-
cule, L
ext
remains constant at first, as the addition of twist is
absorbed through deformation of the molecule (Fig. 6). In this
regime, Tw > Tw
0
,Wr = 0, and the torque Γ increase linearly with
the number of turns N:
Γ=C2πN=LC,ð21Þ
where C is the torsional modulus of the molecule, e.g., C~90 k
B
T
for DNA [35, 7375]. This may be used to monitor the torque-
dependence of a specific DNA-protein interaction, e.g., RNA
polymerase-promoter open complex formation by the bacterial
RNA polymerase [76]. At the critical torque Γ
C
, the molecule’s
extension suddenly decreases to form the first loop upon further
addition of coiling to the DNA molecule (Fig. 6). This event is also
called the buckling transition and is followed by a linear decrease in
L
ext
with added turns [6, 77]. Γ
C
is given through having the
energy to form a loop of radius R
L
being equal to the work done
by the addition of one extra turn 2πΓ
C
:
ER =2πRLFmag þ πLPkBT=RL,ð22Þ
Minimizing E
R
as a function of R
L
gives Γ
C
and the change in
extension per superhelical turn Δz such as [73, 78]:
ΓC = 2LPkBTFmag andΔz =2πRL =π 2LPkBT=Fmag, ð23Þ
This model only describes Δz qualitatively, and more sophisti-
cated models have been derived to describe Δz more accurately
[15, 31]. At high force and in the negative supercoil regime, it is
more favorable for the DNA molecule to unwind than forming
plectoneme, while the rotation-extension of a DNA molecule is
symmetrical at low force, i.e., the tether forms plectonemes for
both negative and positive supercoil addition (Fig. 6).
4 Combining Magnetic Tweezers with Other Techniques
Magnetic tweezers have been combined with fluorescence micros-
copy to enable simultaneous force/torque and fluorescence spec-
troscopy investigations. The preferred fluorescence approach is
objective-based total internal reflection fluorescence microscopy
(TIRFM), as it is a surface-based approach with a shallow excitation
depth (~hundreds of nanometers), leaving the magnetic bead out
of the excitation volume. Two main configurations have been
reported for such assay, using either a standard vertical magnet
configuration (Fig. 7a) or a horizontal magnet configuration
(Fig. 7b). In the former configuration, the magnets pull vertically
on the magnetic bead, and vertical motion of a dye-labeled enzyme
may be reported using the fluorescence channel and the exponen-
tial decay of the evanescent field of the TIRF excitation [7982]. In
the latter configuration, a magnetic force is applied sideways, which
stretches the DNA molecule laterally, enabling transverse observa-
tion of displacements biomolecular objects (e.g., protein, plecto-
neme) [8385]. These two configurations have different
advantages. The first enables rapid modulation of the applied
torque, as is done in the rotor bead assay developed by Bryant
and colleagues [12]. A sideway configuration gives access to a
higher localization precision of fluorescently labeled biomolecules
moving along, e.g., a DNA tether [8385]. Chapter 22 will discuss
different configurations of magnetic tweezers combined with fluo-
rescence microscopy. Darkfield microscopy has been combined
with magnetic tweezers in the rotor bead assay to use backscattered
light from a gold nanoparticle as a tracker, which provides an
excellent signal-to-noise ratio while minimizing the size of the
object to track, and therefore gives access to a higher measurement
bandwidth [86]. The combination of magnetic tweezers with opti-
cal tweezers has also been reported by Cees Dekker and
colleagues [87].
390 David Dulin
Fig. 7 Magnetic tweezers combined with total internal reflection fluorescence (TIRF) microscop y. (a) Vertically
oriented attractive force and (b) horizontally oriented attractive force. The pink oval indicates a protein of
interest labeled with a fluorescent dye (green sphere). In (a), the exponential decay of the evanescent field is
used to localize the protein along the DNA, while in (b) the position is determined from direct localization in the
plane of observation
An Introduction to Magnetic Tweezers 391
5 Applications of Magnetic Tweezers in Single-Molecule Biophysics
Magnetic tweezers present many advantages to study biomolecules
in vitro one at a time. They are a force clamp technique that enables
force spectroscopy measurements from ~10 fN to ~1 nN [4],
depending on the total amount of magnetic material in the bead.
Because the distance between the magnetic bead and the magnets
has to vary significantly (around 0.05 mm) to vary the force signifi-
cantly, magnetic tweezers can apply a constant force over very long
measurement. This holds true even at low force (< 1 pN), unlike
for an AFM or optical tweezers. Furthermore, the combination of a
homogenous magnetic field over a very large field of view (~mm
2
)
and commercially available magnetic beads with homogenous mag-
netic content enables high-throughput force spectroscopy mea-
surements at constant force with a small bead-to-bead variation in
force (~10% standard deviation). For these reasons, magnetic twee-
zers have been applied to study protein folding and unfolding
dynamics at a low constant force [19], such as the titin immuno-
globulin domain [23], talin and protein L [21], von Willebrand
factor folding [20], protein-ligand interactions to interrogate
either SARS-CoV-2 spike or ACE2 interactions (Fig. 8a, b)[88],
and the rapamycin-mediated association between FKBP12 and
FRB [22, 89].
Furthermore, the recent advances in tracking algorithms, illu-
mination, and imaging strategies have brought magnetic tweezers
on par with optical tweezers in terms of their spatiotemporal reso-
lution. Additionally, the ability to perform high-throughput track-
ing in magnetic tweezers enables the in-depth characterization of
mechanochemical pathways of translocating molecular motor.
Examples include viral RNA polymerases (Fig. 8c, d)[26, 49,
9092], the bacterial RNA polymerase [93], the DNA polymerase
[7, 8], helicases (Fig. 5c)[39, 44, 94104], and the SMC complex
[105, 106]. Magnetic tweezers have also enabled the characteriza-
tion of nucleoprotein filament formation or mechanical properties
[107110], protein-mediated DNA condensation [111, 112], and
chromatin filament and nucleosome stability [113117].
Magnetic tweezers are naturally well suited to perform torque
spectroscopy experiments. This has been extensively used to inves-
tigate biological systems that induce a change in the linking number
of a tethered coilable double-stranded nucleic acid. Topoisome-
rases remove the excess of negative or positive supercoils in the
DNA molecule that naturally occur in the cell during DNA tran-
scription and replication [118]. Therefore, their activity is essential
to maintain cellular homeostasis. Using magnetic tweezers to
induce a large excess of supercoils to a DNA molecule has been
used to investigate the mechanochemical cycle of topoisomerases
[24, 119124]. Cellular RNA polymerases have been extensively
392 David Dulin
Fig. 8 Examples of magnetic tweezers applications in single-molecule biophysics. (a, b) Schematic and
experimental traces showing the binding and dissociation kinetics of the SARS-CoV-2 spike protein RBD from
the ACE2 receptor as a function of force. (b) The time-dependent traces reveal populations in the bound and
dissociated states as a function of the applied force (Adapted from Ref. [88]). (c, d) Schematic and
experimental traces of elongating SARS-CoV-2 core replication-transcription complexes. (d) The time-
dependent traces demonstrate rich dynamics with bursts of nucleotide addition interrupted by pauses of
various durations (Adapted from Ref. [49]). (e, f) RNA polymerase (RNAP)-promoter open complex formation on
a positively supercoiled DNA. Upon promoter opening, upon n positive supercoils addition, moving the bead
downward by nΔz. The surface is passivated using a lipid-bilayer strategy. The trace in (f) shows the promoter
alternating between a closed state (CS, promoter closed) and an open state (OS, RNAP-promoter open)
(Adapted from Ref. [50])
investigated using magnetic tweezers. Indeed, they must open
double-stranded DNA at the promoter site to form the RNA
polymerase-promoter open complex and initiate transcription,
which consequently removes ~1 turn of twist in the DNA molecule
(the transcription bubble is 1314 nt, and DNA makes one full turn
every ~10.5 bp). Using the conservation of the linking number
condition described above (Fig. 8e), many details in the mechanism
of transcription initiation by cellular RNA polymerase have been
revealed (Fig. 8f), such as the impact of torque on promoter open-
ing, the dynamics of transcription initiation and promoter escape as
a function of the promoter sequence and salt concentration, the
transcription start site selection, and R-loop formation during
transcription [14, 50, 76, 125129]. A similar approach was used
to investigate Cas9 R-loop formation [130, 131]. The torque
spectroscopy capabilities of magnetic tweezers have also been
used to investigate the torsional properties and stability of chroma-
tin filaments [116, 132, 133] and nucleosome assembly [134
136], as well as other nucleoprotein filaments, such as those formed
by Rad51, RecA, and H-NS [35, 137141].
An Introduction to Magnetic Tweezers 393
6 Perspectives
The unique advantages of magnetic tweezers, i.e., simplicity (and
therefore low cost), stability, high parallelization and resolution,
large force range, and torque spectroscopy, make it a very powerful
technique. It has been established in many labs worldwide, with an
ever-increasing demand. Only one company currently sells mag-
netic tweezers instruments (Mad City labs), and more could be
done to have the technique more available at low cost. Currently,
no open-source instrument design has been released or published,
and CAD drawings would help democratizing magnetic tweezers.
A software interface already exists [25], though in proprietary
format (LabView, National Instruments), and efforts must be
made to release an open-source interface in a nonproprietary lan-
guage, such as Python. The development of routines for data
analysis in nonproprietary languages to help the analysis of complex
dynamics of molecular motors will further support the democrati-
zation of magnetic tweezers with appropriate statistical tools to
analyze single-molecule data. Altogether, these developments will
bring magnetic tweezers and their application to a broader com-
munity. Lastly, there is no combined high-throughput single-
molecule force/torque and fluorescence spectroscopy assay avail-
able to date. The statistical power of such hydrid assay would
potentiate the investigation of ever more complex biomolecular
systems.
394 David Dulin
References
1. Crick FHC, Hughes AFW (1950) The physi-
cal properties of cytoplasm: a study by means
of the magnetic particle method part I. Exp
Cell Res 1(1): 3780. https://doi.org/10.
1016/0014-4827(50)90048-6
2. Kah D, Durrbeck C, Schneider W, Fabry B,
Gerum RC (2020) High-force magnetic
tweezers with hysteresis-free force feedback.
Biophys J 119(1):1523. https://doi.org/
10.1016/j.bpj.2020.05.018
3. Kollmannsberger P, Fabry B (2007) BaHigh-
force magnetic tweezers with force feedback
for biological applications. Rev Sci Instrum
78(11):114301. https://doi.org/10.1063/
1.2804771
4. Dulin D, Lipfert J, Moolman MC, Dekker
NH (2013) Studying genomic processes at
the single-molecule level: introducing the
tools and applications. Nat Rev Genet 14(1):
922. https://doi.org/10.1038/nrg3316
5. Smith SB, Finzi L, Bustamante C (1992)
Direct mechanical measurements of the elas-
ticity of single DNA molecules by using mag-
netic beads. Science (New York, NY)
258(5085):11221126
6. Strick TR, Allemand JF, Bensimon D,
Bensimon A, Croquette V (1996) The elastic-
ity of a single supercoiled DNA molecule.
Science (New York, NY) 271(5257):
18351837
7. Manosas M, Spiering MM, Ding F,
Croquette V, Benkovic SJ (2012) Collabora-
tive coupling between polymerase and heli-
case for leading-strand synthesis. Nucleic
Acids Res 40(13):61876198. https://doi.
org/10.1093/nar/gks254
8. Manosas M, Spiering MM, Ding F,
Bensimon D, Allemand JF, Benkovic SJ, Cro-
quette V (2012) Mechanism of strand dis-
placement synthesis by DNA replicative
polymerases. Nucleic Acids Res 40(13):
61746186. https://doi.org/10.1093/nar/
gks253
9. Maier B, Bensimon D, Croquette V (2000)
Replication by a single DNA polymerase of a
stretched single-stranded DNA. Proc Natl
Acad Sci U S A 97(22):1200212007.
https://doi.org/10.1073/pnas.97.22.
12002
10. Charvin G, Strick TR, Bensimon D, Cro-
quette V (2005) Tracking topoisomerase
activity at the single-molecule level. Annu
Rev Biophys Biomol Struct 34:201219.
https://doi.org/10.1146/annurev.biophys.
34.040204.144433
11. Spakman D, Bakx JAM, Biebricher Andreas S,
Peterman EJG, Wuite GJL, King Graeme A
(2021) Unravelling the mechanisms of Type
1A topoisomerases using single-molecule
approaches. Nucleic Acids Res 49(10):
54705492. https://doi.org/10.1093/nar/
gkab239
12. Basu A, Hobson M, Lebel P, Fernandes LE,
Tretter EM, Berger JM, Bryant Z (2018)
Dynamic coupling between conformations
and nucleotide states in DNA gyrase. Nat
Chem Biol 14(6):565574. https://doi.org/
10.1038/s41589-018-0037-0
13. McKie SJ, Neuman KC, Maxwell A (2021)
DNA topoisomerases: advances in under-
standing of cellular roles and multi-protein
complexes via structure-function analysis.
BioEssays 43(4):e2000286. https://doi.org/
10.1002/bies.202000286
14. Ostrofet E, Papini FS, Malinen AM, Dulin D
(2019) A single-molecule view on cellular and
viral RNA synthesis. In: Joo C, Rueda D (eds)
Biophysics of RNA-protein interactions,
Biological and Medical Physics, Biomedical
Engineering. Springer, New York, pp
109141. https://doi.org/10.1007/978-1-
4939-9726-8
15. Kriegel F, Ermann N, Lipfert J (2017) Prob-
ing the mechanical properties, conformational
changes, and interactions of nucleic acids with
magnetic tweezers. J Struct Biol 197(1):
2636. https://doi.org/10.1016/j.jsb.
2016.06.022
16. De Vlaminck I, Dekker C (2012) Recent
advances in magnetic tweezers. Annu Rev
Biophys 41:453472. https://doi.org/10.
1146/annurev-biophys-122311-100544
17. Morati F, Modesti M (2021) Insights into the
control of RAD51 nucleoprotein filament
dynamics from single-molecule studies. Curr
Opin Genet Dev 71:182187. https://doi.
org/10.1016/j.gde.2021.09.001
18. Winardhi RS, Yan J (2017) Applications of
magnetic tweezers to studies of NAPs. Meth-
ods Mol Biol 1624:173191. https://doi.
org/10.1007/978-1-4939-7098-8_14
19. Zhao X, Zeng X, Lu C, Yan J (2017) Studying
the mechanical responses of proteins using
magnetic tweezers. Nanotechnology 28(41):
414002. https://doi.org/10.1088/1361-
6528/aa837e
20. Lof A, Walker PU, Sedlak SM, Gr uber S,
Obser T, Brehm MA, Benoit M, Lipfert J
(2019) Multiplexed protein force spectros-
copy reveals equilibrium protein folding
An Introduction to Magnetic Tweezers 395
dynamics and the low-force response of von
Willebrand factor. Proc Natl Acad Sci U S A
116(38):1879818807. https://doi.org/10.
1073/pnas.1901794116
21. Tapia-Rojo R, Eckels EC, Fernandez JM
(2019) Ephemeral states in protein folding
under force captured with a magnetic twee-
zers design. Proc Natl Acad Sci U S A
116(16):78737878. https://doi.org/10.
1073/pnas.1821284116
22. Kostrz D, Wayment-Steele HK, Wang JL,
Follenfant M, Pande VS, Strick TR, Gosse C
(2019) A modular DNA scaffold to study
protein-protein interactions at single-
molecule resolution. Nat Nanotechnol
14(10):988993. https://doi.org/10.1038/
s41565-019-0542-7
23. Chen H, Yuan G, Winardhi RS, Yao M,
Popa I, Fernandez JM, Yan J (2015) Dynam-
ics of equilibrium folding and unfolding tran-
sitions of titin immunoglobulin domain under
constant forces. J Am Chem Soc 137(10):
35403546. https://doi.org/10.1021/
ja5119368
24. Agarwal R, Duderstadt KE (2020) Multiplex
flow magnetic tweezers reveal rare enzymatic
events with single molecule precision. Nat
Commun 11(1):4714. https://doi.org/10.
1038/s41467-020-18456-y
25. Cnossen JP, Dulin D, Dekker NH (2014) An
optimized software framework for real-time,
high-throughput tracking of spherical beads.
Rev Sci Instrum 85(10):103712. https://doi.
org/10.1063/1.4898178
26. Dulin D, Vilfan ID, Berghuis BA, Hage S,
Bamford DH, Poranen MM, Depken M,
Dekker NH (2015) Elongation-competent
pauses govern the fidelity of a viral
RNA-dependent RNA polymerase. Cell Rep
10(6):983992. https://doi.org/10.1016/j.
celrep.2015.01.031
27. Berghuis BA, Dulin D, Xu ZQ, van Laar T,
Cross B, Janissen R, Jergic S, Dixon NE,
Depken M, Dekker NH (2015) Strand sepa-
ration establishes a sustained lock at the
Tus-Ter replication fork barrier. Nat Chem
Biol 11(8):579585. https://doi.org/10.
1038/nchembio.1857
28. Dulin D, Cui TJ, Cnossen J, Docter MW,
Lipfert J, Dekker NH (2015) High
spatiotemporal-resolution magnetic tweezers:
calibration and applications for DNA dynam-
ics. Biophys J 109(10):21132125. https://
doi.org/10.1016/j.bpj.2015.10.018
29. Lansdorp BM, Tabrizi SJ, Dittmore A, Saleh
OA (2013) A high-speed magnetic tweezer
beyond 10,000 frames per second. Rev Sci
Instrum 84(4):044301. https://doi.org/10.
1063/1.4802678
30. Huhle A, Klaue D, Brutzer H, Daldrop P,
Joo S, Otto O, Keyser UF, Seidel R (2015)
Camera-based three-dimensional real-time
particle tracking at kHz rates and Angstrom
accuracy. Nat Commun 6:5885. https://doi.
org/10.1038/ncomms6885
31. Vilfan ID, Lipfert J, Koster DA, Lemay SG,
Dekker NH (2009) Magnetic tweezers for
single-molecule experiments. In: Handbook
of single-molecule biophysics, pp 371395.
https://doi.org/10.1007/978-0-387-
76497-9_13
32. Lipfert J, Hao X, Dekker NH (2009) Quanti-
tative modeling and optimization of magnetic
tweezers. Biophys J 96(12):50405049.
https://doi.org/10.1016/j.bpj.2009.
03.055
33. De Vlaminck I, Henighan T, van Loenhout
MT, Burnham DR, Dekker C (2012) Mag-
netic forces and DNA mechanics in multi-
plexed magnetic tweezers. PLoS One 7(8):
e41432. https://doi.org/10.1371/journal.
pone.0041432
34. Lipfert J, Wiggin M, Kerssemakers JWJ,
Pedaci F, Dekker NH (2011) Freely orbiting
magnetic tweezers to directly monitor
changes in the twist of nucleic acids. Nat
Commun 2. doi:Artn 439. https://doi.org/
10.1038/Ncomms1450
35. Lipfert J, Kerssemakers JW, Jager T, Dekker
NH (2010) Magnetic torque tweezers: mea-
suring torsional stiffness in DNA and RecA-
DNA filaments. Nat Methods 7(12):977980
36. Kriegel F, Ermann N, Forbes R, Dulin D,
Dekker NH, Lipfert J (2017) Probing the
salt dependence of the torsional stiffness of
DNA by multiplexed magnetic torque twee-
zers. Nucleic Acids Res 45(10):59205929.
https://doi.org/10.1093/nar/gkx280
37. Otto O, Czerwinski F, Gornall JL, Stober G,
Oddershede LB, Seidel R, Keyser UF (2010)
Real-time particle tracking at 10,000 fps using
optical fiber illumination. Opt Express
18(22):2272222733
38. Dulin D, Barland S, Hachair X, Pedaci F
(2014) Efficient illumination for microsecond
tracking microscopy. PLoS One 9(9):
e107335. https://doi.org/10.1371/journal.
pone.0107335
39. Rieu M, Vieille T, Radou G, Jeanneret R,
Ruiz-Gutierrez N, Ducos B, Allemand JF,
Croquette V (2021) Parallel, linear, and sub-
nanometric 3D tracking of microparticles
with Stereo Darkfield Interferometry. Sci
396 David Dulin
Adv 7(6). https://doi.org/10.1126/sciadv.
abe3902
40. Gosse C, Croquette V (2002) Magnetic twee-
zers: micromanipulation and force measure-
ment at the molecular level. Biophys J 82(6):
33143329
41. Strick T (1999) Mechanical supercoiling of
DNA and its relaxation by topoisomerases.
Ph.D. Thesis, Paris VI, Paris
42. van Loenhout MT, Kerssemakers JW, De
Vlaminck I, Dekker C (2012) Non-bias-lim-
ited tracking of spherical particles, enabling
nanometer resolution at low magnification.
Biophys J 102(10):23622371. https://doi.
org/10.1016/j.bpj.2012.03.073
43. Winzor DJ, Jackson CM (2006) Interpreta-
tion of the temperature dependence of equi-
librium and rate constants. J Mol Recognit
19(5):389407. https://doi.org/10.1002/
jmr.799
44. Gollnick B, Carrasco C, Zuttion F, Gilhooly
NS, Dillingham MS, Moreno-Herrero F
(2015) Probing DNA helicase kinetics with
temperature-controlled magnetic tweezers.
Small 11(11):12731284. https://doi.org/
10.1002/smll.201402686
45. Kriegel F, Matek C, Drsata T, Kulenkampff K,
Tschirpke S, Zacharias M, Lankas F, Lipfert J
(2018) The temperature dependence of the
helical twist of DNA. Nucleic Acids Res
46(15):79988009. https://doi.org/10.
1093/nar/gky599
46. Seifert M, van Nies P, Papini FS, Arnold JJ,
Poranen MM, Cameron CE, Depken M,
Dulin D (2020) Temperature controlled
high-throughput magnetic tweezers show
striking difference in activation energies of
replicating viral RNA-dependent RNA poly-
merases. Nucleic Acids Res 48(10):
55915602. https://doi.org/10.1093/nar/
gkaa233
47. Janissen R, Berghuis BA, Dulin D, Wink M,
van Laar T, Dekker NH (2014) Invincible
DNA tethers: covalent DNA anchoring for
enhanced temporal and force stability in mag-
netic tweezers experiments. Nucleic Acids Res
42(18):e137. https://doi.org/10.1093/
nar/gku677
48. Eeftens JM, van der Torre J, Burnham DR,
Dekker C (2015) Copper-free click chemistry
for attachment of biomolecules in magnetic
tweezers. BMC Biophys 8:9. https://doi.
org/10.1186/s13628-015-0023-9
49. Bera SC, Seifert M, Kirchdoerfer RN, van
Nies P, Wubulikasimu Y, Quack S, Papini FS,
Arnold JJ, Canard B, Cameron CE,
Depken M, Dulin D (2021) The nucleotide
addition cycle of the SARS-CoV-2 polymer-
ase. Cell Rep 36(9):109650. https://doi.
org/10.1016/j.celrep.2021.109650
50. Bera SC, America PPB, Maatsola S, Seifert M,
Ostrofet E, Cnossen J, Spermann M, Papini
FS, Depken M, Malinen AM, Dulin D (2022)
Quantitative parameters of bacterial RNA
polymerase open-complex formation, stabili-
zation and disruption on a consensus pro-
moter. Nucleic Acids Res 50(13):
75117528. https://doi.org/10.1093/nar/
gkac560
51. Papini FS, Seifert M, Dulin D (2019) High-
yield fabrication of DNA and RNA constructs
for single molecule force and torque spectros-
copy experiments. Nucleic Acids Res 47(22):
e144. https://doi.org/10.1093/nar/
gkz851
52. Bell NAW, Molloy JE (2022) Efficient golden
gate assembly of DNA constructs for single
molecule force spectroscopy and imaging.
Nucleic Acids Res 50(13):e77. https://doi.
org/10.1093/nar/gkac300
53. Lu Y, Bianco P (2021) High-yield purification
of exceptional-quality, single-molecule DNA
substrates. J Biol Methods 8(1):e145.
https://doi.org/10.14440/jbm.2021.350
54. Berghuis BA, Kober M, van Laar T, Dekker
NH (2016) High-throughput, high-force
probing of DNA-protein interactions with
magnetic tweezers. Methods 105:9098.
https://doi.org/10.1016/j.ymeth.2016.
03.025
55. Manosas M, Meglio A, Spiering MM, Ding F,
Benkovic SJ, Barre FX, Saleh OA, Allemand
JF, Bensimon D, Croquette V (2010) Mag-
netic tweezers for the study of DNA tracking
motors. Methods Enzymol 475:297320.
https://doi.org/10.1016/S0076-6879(10)
75013-8
56. Paik DH, Roskens VA, Perkins TT (2013)
Torsionally constrained DNA for single-
molecule assays: an efficient, ligation-free
method. Nucleic Acids Res 41(19):e179.
https://doi.org/10.1093/nar/gkt699
57. Charvin G, Allemand JF, Strick TR,
Bensimon D, Croquette V (2004)
Twisting DNA: single molecule studies. Con-
temp Phys 45(5):383403. https://doi.org/
10.1080/00107510410001697279
58. van Oene MM, Dickinson LE, Pedaci F,
Kober M, Dulin D, Lipfert J, Dekker NH
(2015) Biological magnetometry: torque on
superparamagnetic beads in magnetic fields.
Phys Rev Lett 114(21):218301. https://doi.
org/10.1103/PhysRevLett.114.218301
An Introduction to Magnetic Tweezers 397
59. Yu Z, Dulin D, Cnossen J, Kober M, van
Oene MM, Ordu O, Berghuis BA,
Hensgens T, Lipfert J, Dekker NH (2014) A
force calibration standard for magnetic twee-
zers. Rev Sci Instrum 85(12):123114.
https://doi.org/10.1063/1.4904148
60. te Velthuis A, Kerssemakers JWJ, Lipfert J,
Dekker NH (2010) Quantitative guidelines
for force calibration through spectral analysis
of magnetic tweezers data. Biophys J 99(4):
12921302. https://doi.org/10.1016/j.bpj.
2010.06.008
61. Ostrofet E, Papini FS, Dulin D (2018)
Correction-free force calibration for magnetic
tweezers experiments. Sci Rep 8(1):15920.
https://doi.org/10.1038/s41598-018-
34360-4
62. Chen H, Fu H, Zhu X, Cong P, Nakamura F,
Yan J (2011) Improved high-force magnetic
tweezers for stretching and refolding of pro-
teins and short DNA. Biophys J 100(2):
517523. https://doi.org/10.1016/j.bpj.
2010.12.3700
63. Schaffer E, Norrelykke SF, Howard J (2007)
Surface forces and drag coefficients of micro-
spheres near a plane surface measured with
optical tweezers. Langmuir 23(7):
36543665. https://doi.org/10.1021/
la0622368
64. Lansdorp BM, Saleh OA (2012) Power spec-
trum and Allan variance methods for calibrat-
ing single-molecule video-tracking
instruments. Rev Sci Instrum 83(2):025115.
https://doi.org/10.1063/1.3687431
65. Morgan IL, Saleh OA (2021) Tweezepy: a
python package for calibrating forces in
single-molecule video-tracking experiments.
PLoS One 16(12):e0262028. https://doi.
org/10.1371/journal.pone.0262028
66. Daldrop P, Brutzer H, Huhle A, Kauert DJ,
Seidel R (2015) Extending the range for force
calibration in magnetic tweezers. Biophys J
108(10):25502561. https://doi.org/10.
1016/j.bpj.2015.04.011
67. Shon MJ, Rah SH, Yoon TY (2019) Submic-
rometer elasticity of double-stranded DNA
revealed by precision force-extension mea-
surements with magnetic tweezers. Sci Adv
5(6):eaav1697. https://doi.org/10.1126/
sciadv.aav1697
68. Sitters G, Kamsma D, Thalhammer G, Ritsch-
Marte M, Peterman EJ, Wuite GJ (2015)
Acoustic force spectroscopy. Nat Methods
12(1):4750. https://doi.org/10.1038/
nmeth.3183
69. Ostrofet E, Papini FS, Dulin D (2020) High
spatiotemporal resolution data from a custom
magnetic tweezers instrument. Data Brief 30:
105397. https://doi.org/10.1016/j.d ib.
2020.105397
70. Bustamante C, Marko JF, Siggia ED, Smith S
(1994) Entropic elasticity of lambda-phage
DNA. Science (New York, NY) 265(5178):
15991600
71. Burnham DR, De Vlaminck I, Henighan T,
Dekker C (2014) Skewed brownian fluctua-
tions in single-molecule magnetic tweezers.
PLoS One 9(9):e108271. https://doi.org/
10.1371/journal.pone.0108271
72. Kriegel F, Vanderlinden W, Nicolaus T,
Kardinal A, Lipfert J (2018) Measuring
single-molecule twist and torque in multi-
plexed magnetic tweezers. Methods Mol Biol
1814:7598. https://doi.org/10.1007/
978-1-4939-8591-3_6
73. Strick TR, Allemand JF, Bensimon D, Cro-
quette V (2000) Stress-induced structural
transitions in DNA and proteins. Annu Rev
Biophys Biomol Struct 29:523543. https://
doi.org/10.1146/annurev.biophys.29.1.523
74. Forth S, Deufel C, Sheinin MY, Daniels B,
Sethna JP, Wang MD (2008) Abrupt buckling
transition observed during the plectoneme
formation of individual DNA molecules.
Phys Rev Lett 100(14):148301. https:/
/doi.
org/10.1103/PhysRevLett.100.148301
75. Bryant Z, Stone MD, Gore J, Smith SB, Coz-
zarelli NR, Bustamante C (2003) Structural
transitions and elasticity from torque mea-
surements on DNA. Nature 424(6946):
338341. https://doi.org/10.1038/
nature01810. nature01810 [pii]
76. Revyakin A, Ebright RH, Strick TR (2004)
Promoter unwinding and promoter clearance
by RNA polymerase: detection by single-
molecule DNA nanomanipulation. Proc Natl
Acad Sci U S A 101(14):47764780. https://
doi.org/10.1073/pnas.0307241101
77. Brutzer H, Luzzietti N, Klaue D, Seidel R
(2010) Energetics at the DNA supercoiling
transition. Biophys J 98(7):12671276.
https://doi.org/10.1016/j.bpj.2009.12.
4292
78. Strick TR, Dessinges MN, Charvin G, Dekker
NH, Allemand JF, Bensimon D, Croquette V
(2003) Stretching of macromolecules and
proteins. Rep Prog Phys 66(1):145
79. Graves ET, Duboc C, Fan J, Stransky F,
Leroux-Coyau M, Strick TR (2015) A
dynamic DNA-repair complex observed by
correlative single-molecule nanomanipulation
and fluorescence. Nat Struct Mol Biol 22(6):
452457. https://doi.org /10.103 8/nsmb.
3019
398 David Dulin
80. Seol Y, Neuman KC (2018) Combined mag-
netic tweezers and micro-mirror total internal
reflection fluorescence microscope for single-
molecule manipulation and visualization.
Methods Mol Biol 1665:297316. https://
doi.org/10.1007/978-1-4939-7271-5_16
81. Brutzer H, Schwarz FW, Seidel R (2012)
Scanning evanescent fields using a pointlike
light source and a nanomechanical DNA
gear. Nano Lett 12(1):473478. https://doi.
org/10.1021/nl203876w
82. Kemmerich FE, Swoboda M, Kauert DJ,
Grieb MS, Hahn S, Schwarz FW, Seidel R,
Schlierf M (2016) Simultaneous single-
molecule force and fluorescence sampling of
DNA nanostructure conformations using
magnetic tweezers. Nano Lett 16(1):
381386. https://doi.org/10.1021/acs.
nanolett.5b03956
83. van Loenhout MT, de Grunt MV, Dekker C
(2012) Dynamics of DNA supercoils. Science
(New York, NY) 338(6103):9497. https://
doi.org/10.1126/science.1225810
84. Schwarz FW, Toth J, van Aelst K, Cui G,
Clausing S, Szczelkun MD, Seidel R (2013)
The helicase-like domains of type III restric-
tion enzymes trigger long-range diffusion
along DNA. Science (New York, NY)
340(6130):353356. https://doi.org/10.
1126/science.1231122
85. Madariaga-Marcos J, Hormeno S, Pastrana
CL, Fisher GLM, Dillingham MS, Moreno-
Herrero F (2018) Force determination in lat-
eral magnetic tweezers combined with TIRF
microscopy. Nanoscale 10(9):45794590.
https://doi.org/10.1039/c7nr07344e
86. Lebel P, Basu A, Oberstrass FC, Tretter EM,
Bryant Z (2014) Gold rotor bead tracking for
high-speed measurements of DNA twist,
torque and extension. Nat Methods 11(4):
456462. https://doi.org/10.1038/nmeth.
2854
87. De Vlaminck I, van Loenhout MT, Zweifel L,
den Blanken J, Hooning K, Hage S,
Kerssemakers J, Dekker C (2012) Mechanism
of homology recognition in DNA recombina-
tion from dual-molecule experiments. Mol
Cell 46(5):616624. https://doi.org/10.
1016/j.molcel.2012.03.029
88. Bauer MS, Gruber S, Hausch A, Gomes P,
Milles LF, Nicolaus T, Schendel LC, Navajas
PL, Procko E, Lietha D, Melo MCR, Bernardi
RC, Gaub HE, Lipfert J (2022) A tethered
ligand assay to probe SARS-CoV-2:ACE2
interactions. Proc Natl Acad Sci U S A
119(14):e2114397119. https://doi.org/10.
1073/pnas.2114397119
89. Wang Y, Barnett SFH, Le S, Guo Z, Zhong X,
Kanchanawong P, Yan J (2019) Label-free
single-molecule quantification of rapamycin-
induced FKBP-FRB dimerization for direct
control of cellular mechanotransduction.
Nano Lett 19(10):75147525. https://doi.
org/10.1021/acs.nanolett.9b03364
90. Seifert M, Bera SC, van Nies P, Kirchdoerfer
RN, Shannon A, Le TT, Meng X, Xia H,
Wood JM, Harris LD, Papini FS, Arnold JJ,
Almo S, Grove TL, Shi PY, Xiang Y,
Canard B, Depken M, Cameron CE, Dulin
D (2021) Inhibition of SARS-CoV-2 poly-
merase by nucleotide analogs from a single-
molecule perspective. elife 10. https://doi.
org/10.7554/eLife.70968
91. Dulin D, Arnold JJ, van Laar T, Oh HS,
Lee C, Perkins AL, Harki DA, Depken M,
Cameron CE, Dekker NH (2017) Signatures
of nucleotide analog incorporation by an
RNA-dependent RNA polymerase revealed
using high-throughput magnetic tweezers.
Cell Rep 21(4):10631076. https://doi.
org/10.1016/j.celrep.2017.10.005
92. Dulin D, Vilfan ID, Berghuis BA, Poranen
MM, Depken M, Dekker NH (2015) Back-
tracking behavior in viral RNA-dependent
RNA polymerase provides the basis for a sec-
ond initiation site. Nucleic Acids Res 43(21):
1042110429. https://doi.org/10.1093/
nar/gkv1098
93. Janissen R, Eslami-Mossallam B,
Artsimovitch I, Depken M, Dekker NH
(2022) High-throughput single-molecule
experiments reveal heterogeneity, state
switching, and three interconnected pause
states in transcription. Cell Rep 39(4):
110749. https://doi.org/10.1016/j.celrep.
2022.110749
94. Hodeib S, Raj S, Manosas M, Zhang W,
Bagchi D, Ducos B, Allemand JF,
Bensimon D, Croquette V (2016) Single mol-
ecule studies of helicases with magnetic twee-
zers. Methods 105:315. https://doi.org/
10.1016/j.ymeth.2016.06.019
95. Rieu M, Valle-Orero J, Ducos B, Allemand JF,
Croquette V (2021) Single-molecule kinetic
locking allows fluorescence-free quantifica-
tion of protein/nucleic-acid binding. Com-
mun Biol 4(1):1083. https://doi.org/10.
1038/s42003-021-02606-z
96. Burnham DR, Kose HB, Hoyle RB, Yardimci
H (2019) The mechanism of DNA unwind-
ing by the eukaryotic replicative helicase. Nat
Commun 10(1):2159. https://doi.org/10.
1038/s41467-019-09896-2
97. Seol Y, Harami GM, Kovacs M, Neuman KC
(2019) Homology sensing via non-linear
An Introduction to Magnetic Tweezers 399
amplification of sequence-dependent pausing
by RecQ helicase. elife 8. https://doi.org/10.
7554/eLife.45909
98. Mills M, Harami GM, Seol Y, Gyimesi M,
Martina M, Kovacs ZJ, Kovacs M, Neuman
KC (2017) RecQ helicase triggers a binding
mode change in the SSB-DNA complex to
efficiently initiate DNA unwinding. Nucleic
Acids R es 45(20):1187811890. https://
doi.org/10.1093/nar/gkx939
99. Klaue D, Kobbe D, Kemmerich F,
Kozikowska A, Puchta H, Seidel R (2013)
Fork sensing and strand switching control
antagonistic activities of RecQ helicases. Nat
Commun 4:2024. https://doi.org/10.
1038/ncomms3024
100. Lionnet T, Spiering MM, Benkovic SJ,
Bensimon D, Croquette V (2007) Real-time
observation of bacteriophage T4 gp41 heli-
case reveals an unwinding mechanism. Proc
Natl Acad Sci U S A 104(50):1979019795.
ht tps:// do i.o r g/1 0 .10 7 3/ pna s .
0709793104. 0709793104 [pii]
101. Dessinges MN, Lionnet T, Xi XG,
Bensimon D, Croquette V (2004) Single-
molecule assay reveals strand switching and
enhanced processivity of UvrD. Proc Natl
Acad Sci U S A 101(17):64396444.
ht tps:// do i.o r g/1 0 .10 7 3/ pna s .
0306713101. 0306713101 [pii]
102. Manosas M, Xi XG, Bensimon D, Croquette
V (2010) Active and passive mechanisms of
helicases. Nucleic Acids Res 38(16):
55185526. https://doi.org/10.1093/nar/
gkq273
103. Fiorini F, Bagchi D, Le Hir H, Croquette V
(2015) Human Upf1 is a highly processive
RNA helicase and translocase with RNP
remodelling activities. Nat Commun 6:
7581. https://doi.org/10.1038/
ncomms8581
104. Bagchi D, Manosas M, Zhang W, Manthei
KA, Hodeib S, Ducos B, Keck JL, Croquette
V (2018) Single molecule kinetics uncover
roles for E. coli RecQ DNA helicase domains
and interaction with SSB. Nucleic Acids Res
46(16):85008515. https://doi.org/10.
1093/nar/gky647
105. Eeftens JM, Bisht S, Kerssemakers J,
Kschonsak M, Haering CH, Dekker C
(2017) Real-time detection of condensin-
driven DNA compaction reveals a multistep
binding mechanism. EMBO J 36(23):
34483457. https://doi.org/10.15252/
embj.201797596
106. Elbatsh AMO, Kim E, Eeftens JM, Raaij-
makers JA, van der Weide RH, Garcia-Nieto-
A, Bravo S, Ganji M, Uit de Bos J,
Teunissen H, Medema RH, de Wit E, Haer-
ing CH, Dekker C, Rowland BD (2019) Dis-
tinct roles for condensin’s two ATPase sites in
chromosome condensation. Mol Cell 76(5):
724737. e725. https://doi.org/10.1016/j.
molcel.2019.09.020
107. van der Heijden T, Seidel R, Modesti M,
Kanaar R, Wyman C, Dekker C (2007) Real-
time assembly and disassembly of human
RAD51 filaments on individual DNA mole-
cules. Nucleic Acids Res 35(17):56465657.
https://doi.org/10.1093/nar/gkm629
108. Liu Y, Chen H, Kenney LJ, Yan J (2010) A
divalent switch drives H-NS/DNA-binding
conformations between stiffening and bridg-
ing modes. Genes Dev 24(4):339344.
https://doi.org/10.1101/gad.1883510
109. Gulvady R, Gao Y, Kenney LJ, Yan J (2018) A
single molecule analysis of H-NS uncouples
DNA binding affinity from DNA specificity.
Nucleic Acids Res 46(19):1021610224.
https://doi.org/10.1093/nar/gky826
110. Fu H, Le S, Muniyappa K, Yan J (2013)
Dynamics and regulation of RecA polymeri-
zation and de-polymerization on double-
stranded DNA. PLoS One 8(6):e66712.
https://doi.org/10.1371/journal.pone.
0066712
111. Vtyurina NN, Dulin D, Docter MW, Meyer
AS, Dekker NH, Abbondanzieri EA (2016)
Hysteresis in DNA compaction by Dps is
described by an Ising model. Proc Natl Acad
Sci U S A 113(18):49824987. https://doi.
org/10.1073/pnas.1521241113
112. Bell NAW, Haynes PJ, Brunner K, de Oliveira
TM, Flocco MM, Hoogenboom BW, Molloy
JE (2021) Single-molecule measurements
reveal that PARP1 condenses DNA by loop
stabilization. Sci Adv 7(33). https://doi.org/
10.1126/sciadv.abf3641
113. Meng H, Andresen K, van Noort J (2015)
Quantitative analysis of single-molecule force
spectroscopy on folded chromatin fibers.
Nucleic Acids Res 43(7):35783590.
https://doi.org/10.1093/nar/gkv215
114. Kruithof M, Chien FT, Routh A, Logie C,
Rhodes D, van Noort J (2009) Single-
molecule force spectroscopy reveals a highly
compliant helical folding for the 30-nm chro-
matin fiber. Nat Struct Mol Biol 16(5):
534540. https://doi.org/10.1038/nsmb.
1590
115. Kruithof M, Chien F, de Jager M, van Noort J
(2008) Subpiconewton dynamic force spec-
troscopy using magnetic tweezers. Biophys J
94(6):23432348. biophysj.107.121673
[pii]. https://doi.org/10.1529/biophysj.
107.121673
400 David Dulin
116. Bancaud A, Conde e Silva N, Barbi M, Wag-
ner G, Allemand JF, Mozziconacci J,
Lavelle C, Croquette V, Victor JM,
Prunell A, Viovy JL (2006) Structural plastic-
ity of single chromatin fibers revealed by tor-
sional manipulation. Nat Struct Mol Biol
13(5):444450. https://doi.org/10.1038/
nsmb1087. doi:nsmb1087 [pii]
117. Simon M, North JA, Shimko JC, Forties RA,
Ferdinand MB, Manohar M, Zhang M,
Fishel R, Ottesen JJ, Poirier MG (2011) His-
tone fold modifications control nucleosome
unwrapping and disassembly. Proc Natl Acad
Sci U S A 108(31):1271112716. https://
doi.org/10.1073/pnas.1106264108
118. Koster DA, Crut A, Shuman S, Bjornsti MA,
Dekker NH (2010) Cellular strategies for reg-
ulating DNA supercoiling: a single-molecule
perspective. Cell 142(4):519530. https://
doi.org/10.1016/j.cell.2010.08.001.
S0092-8674(10)00896-2 [pii]
119. Lipfert J, Koster DA, Vilfan ID, Hage S, Dek-
ker NH (2009) Single-molecule magnetic
tweezers studies of type IB topoisomerases.
DNA Topoisomerases: Methods and Proto-
cols 582:7189. https://doi.org/10.1007/
978-1-60761-340-4_7
120. Mills M, Tse-Dinh YC, Neuman KC (2018)
Direct observation of topoisomerase IA gate
dynamics. Nat Struct Mol Biol 25(12):
11111118. https://doi.org/1 0.1038/
s41594-018-0158-x
121. McKie SJ, Desai PR, Seol Y, Allen AM,
Maxwell A, Neuman KC (2022) Topoisomer-
ase VI is a chirally-selective, preferential DNA
decatenase. elife 11. https://doi.org/10.
7554/eLife.67021
122. Strick TR, Croquette V, Bensimon D (2000)
Single-molecule analysis of DNA uncoiling by
a type II topoisomerase. Nature 404(6780):
901904. https://doi.org/10.1038/
35009144
123. Koster DA, Palle K, Bot ES, Bjornsti MA,
Dekker NH (2007) Antitumour drugs
impede DNA uncoiling by topoisomerase
I. Nature 448(7150):213217. https://doi.
org/10.1038/nature05938
124. Koster DA, Croquette V, Dekker C,
Shuman S, Dekker NH (2005) Friction and
torque govern the relaxation of DNA super-
coils by eukaryotic topoisomerase IB. Nature
434(7033):671674. https://doi.org/10.
1038/nature03395
125. Revyakin A, Liu C, Ebright RH, Strick TR
(2006) Abortive initiation and productive ini-
tiation by RNA polymerase involve DNA
scrunching. Science (New York, NY)
314(5802):11391143. https://doi.org/10.
1126/science.1131398
126. Tomko EJ, Fishburn J, Hahn S, Galburt EA
(2017) TFIIH generates a six-base-pair open
complex during RNAP II transcription initia-
tion and start-site scanning. Nat Struct Mol
Biol 24(12):11391145. https://doi.org/
10.1038/nsmb.3500
127. Lerner E, Chung S, Allen BL, Wang S, Lee J,
Lu SW, Grimaud LW, Ingargiola A,
Michalet X, Alhadid Y, Borukhov S, Strick
TR, Taatjes DJ, Weiss S (2016) Backtracked
and paused transcription initiation intermedi-
ate of Escherichia coli RNA polymerase. Proc
Natl Acad Sci U S A 113(43):E6562E6571.
ht tps:// do i.o r g/1 0 . 10 7 3 / p na s.
1605038113
128. Yu LB, Winkelman JT, Pukhrambam C, Strick
TR, Nickels BE, Ebright RH (2017) The
mechanism of variability in transcription start
site selection. elife 6. ARTN e32038. https:/
/
doi.org/10.7554/eLife.32038
129. Portman JR, Brouwer GM, Bollins J, Savery
NJ, Strick TR (2021) Cotranscriptional
R-loop formation by Mfd involves topological
partitioning of DNA. Proc Natl Acad Sci U S
A 118(15). https://doi.org/10.1073/pnas.
2019630118
130. Szczelkun MD, Tikhomirova MS,
Sinkunas T, Gasiunas G, Karvelis T,
Pschera P, Siksnys V, Seidel R (2014) Direct
observation of R-loop formation by single
RNA-guided Cas9 and Cascade effector com-
plexes. Proc Natl Acad Sci U S A 111(27):
97989803. https://doi.org/10.1073/pnas.
1402597111
131. Rutkauskas M, Sinkunas T, Songailiene I,
Tikhomirova MS, Siksnys V, Seidel R (2015)
Directional R-loop formation by the
CRISPR-Cas surveillance complex cascade
provides efficient off-target site rejection.
Cell Rep. https://doi.org/10.1016/j.celrep.
2015.01.067
132. Kaczmarczyk A, Meng H, Ordu O, Noort JV,
Dekker NH (2020) Chromatin fibers stabilize
nucleosomes under torsional stress. Nat
Commun 11(1):126. https://doi.org/10.
1038/s41467-019-13891-y
133. Henneman B, Brouwer TB, Erkelens AM,
Kuijntjes GJ, van Emmerik C, van der Valk
RA, Timmer M, Kirolos NCS, van Ingen H,
van Noort J, Dame RT (2021) Mechanical
and structural properties of archaeal hypernu-
cleosomes. Nucleic Acids Res 49(8):
43384349. https://doi.org/10.1093/nar/
gkaa1196
134. Vlijm R, Lee M, Lipfert J, Lusser A,
Dekker C, Dekker NH (2015) Nucleosome
Open Access This chapter is licensed under the term s of the Creative Commons Attribution 4.0 International
License ( ), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as y ou give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons licens e and indicate if changes were made.
The images or other third party material in this chapt er are included in the chapter’s Creative Commons license,
unless indicated other wise in a credit line to the mat erial. If material is not included in the chapter’s Creative
Commons license and your intended use is not permit ted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the c opyright holder.
http://creativecommons.org/licenses/by/4. 0/
An Introduction to Magnetic Tweezers 401
assembly dynamics involve spontaneous fluc-
tuations in the handedness of tetrasomes. Cell
Rep 10(2):216225. https://doi.org/10.
1016/j.celrep.2014.12.022
135. Vlijm R, Kim SH, De Zwart PL, Dalal Y,
Dekker C (2017) The supercoiling state of
DNA determines the handedness of both
H3 and CENP-A nucleosomes. Nanoscale
9(5):18621870. https://doi.org/10.1039/
c6nr06245h
136. Ordu O, Kremser L, Lusser A, Dekker NH
(2018) Modification of the histone tetramer
at the H3-H3 interface impacts tetrasome
conformations and dynamics. J Chem Phys
148(12):123323. https://doi.org/10.
1063/1.5009100
137. Lee M, Lipfert J, Sanchez H, Wyman C, Dek-
ker NH (2013) Structural and torsional prop-
erties of the RAD51-dsDNA nucleoprotein
filament. Nucleic Acids Res 41(14):
70237030. https://doi.org/10.1093/nar/
gkt425
138. Fulconis R, Bancaud A, Allemand JF,
Croquette V, Dutreix M, Viovy JL (2004)
Twisting and untwisting a single DNA mole-
cule covered by RecA protein. Biophys J
87(4):25522563. https://doi.org/10.
1529/biophysj.104.043059
139. Atwell S, Disseau L, Stasiak AZ, Stasiak A,
Renodon-Corniere A, Takahashi M, Viovy
JL, Cappello G (2012) Probing Rad51-
DNA interactions by changing DNA twist.
Nucleic Acids Res 40(22):1176911776.
https://doi.org/10.1093/nar/gks1131
140. Arata H, Dupont A, Mine-Hattab J,
Disseau L, Renodon-Corniere A,
Takahashi M, Viovy JL, Cappello G (2009)
Direct observation of twisting steps during
Rad51 polymerization on DNA. Proc Natl
Acad Sci U S A 106(46):1923919244.
ht tps:// do i.o r g/1 0 . 10 7 3 / p na s.
0902234106. 0902234106 [pii]
141. Lim CJ, Kenney LJ, Yan J (2014) Single-
molecule studies on the mechanical interplay
between DNA supercoiling and H-NS DNA
architectural properties. Nucleic Acids Res
42(13):83698378. https://doi.org/10.
1093/nar/gku566
Article
RNA plays critical roles in the transmission and regulation of genetic information and is increasingly used in biomedical and biotechnological applications. Functional RNAs contain extended double-stranded regions, and the structure of double-stranded RNA (dsRNA) has been revealed at high resolution. However, the dependence of the properties of the RNA double helix on environmental effects, notably temperature, is still poorly understood. Here, we use single-molecule magnetic tweezer measurements to determine the dependence of the dsRNA twist on temperature. We find that dsRNA unwinds with increasing temperature, even more than DNA, with ΔTwRNA = −14.4 ± 0.7°/(°C·kbp), compared to ΔTwDNA = −11.0 ± 1.2°/(°C·kbp). All-atom molecular dynamics (MD) simulations using a range of nucleic acid force fields, ion parameters, and water models correctly predict that dsRNA unwinds with rising temperature but significantly underestimate the magnitude of the effect. These MD data, together with additional MD simulations involving DNA and DNA–RNA hybrid duplexes, reveal a linear correlation between the twist temperature decrease and the helical rise, in line with DNA but at variance with RNA experimental data. We speculate that this discrepancy might be caused by some unknown bias in the RNA force fields tested or by as yet undiscovered transient alternative structures in the RNA duplex. Our results provide a baseline to model more complex RNA assemblies and to test and develop new parametrizations for RNA simulations. They may also inspire physical models of the temperature-dependent dsRNA structure.
Article
Full-text available
Transcription initiation is the first step in gene expression, and is therefore strongly regulated in all domains of life. The RNA polymerase (RNAP) first associates with the initiation factor $\sigma$ to form a holoenzyme, which binds, bends and opens the promoter in a succession of reversible states. These states are critical for transcription regulation, but remain poorly understood. Here, we addressed the mechanism of open complex formation by monitoring its assembly/disassembly kinetics on individual consensus lacUV5 promoters using high-throughput single-molecule magnetic tweezers. We probed the key protein–DNA interactions governing the open-complex formation and dissociation pathway by modulating the dynamics at different concentrations of monovalent salts and varying temperatures. Consistent with ensemble studies, we observed that RNAP-promoter open (RPO) complex is a stable, slowly reversible state that is preceded by a kinetically significant open intermediate (RPI), from which the holoenzyme dissociates. A strong anion concentration and type dependence indicates that the RPO stabilization may involve sequence-independent interactions between the DNA and the holoenzyme, driven by a non-Coulombic effect consistent with the non-template DNA strand interacting with $\sigma$ and the RNAP $\beta$ subunit. The temperature dependence provides the energy scale of open-complex formation and further supports the existence of additional intermediates.
Article
Full-text available
Single-molecule techniques such as optical tweezers and fluorescence imaging are powerful tools for probing the biophysics of DNA and DNA-protein interactions. The application of these methods requires efficient approaches for creating designed DNA structures with labels for binding to a surface or microscopic beads. In this paper, we develop a simple and fast technique for making a diverse range of such DNA constructs by combining PCR amplicons and synthetic oligonucleotides using golden gate assembly rules. We demonstrate high yield fabrication of torsionally-constrained duplex DNA up to 10 kbp in length and a variety of DNA hairpin structures. We also show how tethering to a cross-linked antibody substrate significantly enhances measurement lifetime under high force. This rapid and adaptable fabrication method streamlines the assembly of DNA constructs for single molecule biophysics.
Article
Full-text available
Significance In the dynamic environment of the airways, where SARS-CoV-2 infections are initiated by binding to human host receptor ACE2, mechanical stability of the viral attachment is a crucial fitness advantage. Using single-molecule force spectroscopy techniques, we mimic the effect of coughing and sneezing, thereby testing the force stability of SARS-CoV-2 RBD:ACE2 interaction under physiological conditions. Our results reveal a higher force stability of SARS-CoV-2 binding to ACE2 compared to SARS-CoV-1, causing a possible fitness advantage. Our assay is sensitive to blocking agents preventing RBD:ACE2 bond formation. It will thus provide a powerful approach to investigate the modes of action of neutralizing antibodies and other agents designed to block RBD binding to ACE2 that are currently developed as potential COVID-19 therapeutics.
Article
Full-text available
DNA topoisomerase VI (topo VI) is a type IIB DNA topoisomerase found predominantly in archaea and some bacteria, but also in plants and algae. Since its discovery, topo VI has been proposed to be a DNA decatenase, however robust evidence and a mechanism for its preferential decatenation activity was lacking. Using single-molecule magnetic tweezers measurements and supporting ensemble biochemistry, we demonstrate that Methanosarcina mazei topo VI preferentially unlinks, or decatenates DNA crossings, in comparison to relaxing supercoils, through a preference for certain DNA crossing geometries. In addition, topo VI demonstrates a significant increase in ATPase activity, DNA binding and rate of strand passage, with increasing DNA writhe, providing further evidence that topo VI is a DNA crossing sensor. Our study strongly suggests that topo VI has evolved an intrinsic preference for the unknotting and decatenation of interlinked chromosomes by sensing and preferentially unlinking DNA crossings with geometries close to 90°.
Article
Full-text available
Single-molecule force spectroscopy (SMFS) instruments (e.g., magnetic and optical tweezers) often use video tracking to measure the three-dimensional position of micron-scale beads under an applied force. The force in these experiments is calibrated by comparing the bead trajectory to a thermal motion-based model with the drag coefficient, γ , and trap spring constant, κ , as parameters. Estimating accurate parameters is complicated by systematic biases from spectral distortions, the camera exposure time, parasitic noise, and least-squares fitting methods. However, while robust calibration methods exist that correct for these biases, they are not always used because they can be complex to implement computationally. To address this barrier, we present Tweezepy: a Python package for calibrating forces in SMFS video-tracking experiments. Tweezepy uses maximum likelihood estimation (MLE) to estimate parameters and their uncertainties from a single bead trajectory via the power spectral density (PSD) and Allan variance (AV). It is well-documented, fast, easy to use, and accounts for most common sources of biases in SMFS video-tracking experiments. Here, we provide a comprehensive overview of Tweezepy’s calibration scheme, including a review of the theory underlying thermal motion-based parameter estimates, a discussion of the PSD, AV, and MLE, and an explanation of their implementation.
Article
Full-text available
The absence of ‘shovel-ready’ anti-coronavirus drugs during vaccine development has exceedingly worsened the SARS-CoV-2 pandemic. Furthermore, new vaccine-resistant variants and coronavirus outbreaks may occur in the near future, and we must be ready to face this possibility. However, efficient antiviral drugs are still lacking to this day, due to our poor understanding of the mode of incorporation and mechanism of action of nucleotides analogs that target the coronavirus polymerase to impair its essential activity. Here, we characterize the impact of remdesivir (RDV, the only FDA-approved anti-coronavirus drug) and other nucleotide analogs (NAs) on RNA synthesis by the coronavirus polymerase using a high-throughput, single-molecule, magnetic-tweezers platform. We reveal that the location of the modification in the ribose or in the base dictates the catalytic pathway(s) used for its incorporation. We show that RDV incorporation does not terminate viral RNA synthesis, but leads the polymerase into backtrack as far as 30 nt, which may appear as termination in traditional ensemble assays. SARS-CoV-2 is able to evade the endogenously synthesized product of the viperin antiviral protein, ddhCTP, though the polymerase incorporates this NA well. This experimental paradigm is essential to the discovery and development of therapeutics targeting viral polymerases.
Article
Full-text available
Genomic integrity depends on the RecA/RAD51 protein family. Discovered over five decades ago with the founder bacterial RecA protein, eukaryotic RAD51 is an ATP-dependent DNA strand transferase implicated in DNA double-strand break and single-strand gap repair, and in dealing with stressed DNA replication forks. RAD51 assembles as a nucleoprotein filament around single-stranded DNA to promote homology recognition in a duplex DNA and subsequent strand exchange. While the intrinsic dynamics of the RAD51 nucleoprotein filament has been extensively studied, a plethora of accessory factors control its dynamics. Understanding how modulators control filament dynamics is at the heart of current research efforts. Here, we describe recent advances in RAD51 control mechanisms obtained specifically using fluorescence-based single-molecule techniques.
Article
Full-text available
Fluorescence-free micro-manipulation of nucleic acids (NA) allows the functional characterization of DNA/RNA processing proteins, without the interference of labels, but currently fails to detect and quantify their binding. To overcome this limitation, we developed a method based on single-molecule force spectroscopy, called kinetic locking, that allows a direct in vitro visualization of protein binding while avoiding any kind of chemical disturbance of the protein’s natural function. We validate kinetic locking by measuring accurately the hybridization energy of ultrashort nucleotides (5, 6, 7 bases) and use it to measure the dynamical interactions of Escherichia coli/E. coli RecQ helicase with its DNA substrate. Rieu et al. present a magnetic tweezers based single-molecule manipulation method, called kinetic locking, for direct detection of biomolecular binding without use of fluorescent probes. By measuring dynamical interactions of E. coli RecQ helicase with its DNA substrate, authors show that this method holds promise for studying DNA-DNA and DNA-protein interactions while avoiding the need for labelling.
Article
Full-text available
Coronaviruses have evolved elaborate multisubunit machines to replicate and transcribe their genomes. Central to these machines are the RNA-dependent RNA polymerase subunit (nsp12) and its intimately associated cofactors (nsp7 and nsp8). We use a high-throughput magnetic-tweezers approach to develop a mechanochemical description of this core polymerase. The core polymerase exists in at least three catalytically distinct conformations, one being kinetically consistent with incorporation of incorrect nucleotides. We provide evidence that an RdRp uses a thermal ratchet instead of a power stroke to transition from the pre- to post-translocated state. Ultra-stable magnetic tweezers enables the direct observation of coronavirus polymerase deep and long-lived backtrack that are strongly stimulated by secondary structure in the template. The framework we present here elucidates one of the most important structure-dynamics-function relationships in human health today, and will form the grounds for understanding the regulation of this complex.
Article
Pausing by bacterial RNA polymerase (RNAp) is vital in the recruitment of regulatory factors, RNA folding, and coupled translation. While backtracking and intra-structural isomerization have been proposed to trigger pausing, our mechanistic understanding of backtrack-associated pauses and catalytic recovery remains incomplete. Using high-throughput magnetic tweezers, we examine the Escherichia coli RNAp transcription dynamics over a wide range of forces and NTP concentrations. Dwell-time analysis and stochastic modeling identify, in addition to a short-lived elemental pause, two distinct long-lived backtrack pause states differing in recovery rates. We identify two stochastic sources of transcription heterogeneity: alterations in short-pause frequency that underlies elongation-rate switching, and variations in RNA cleavage rates in long-lived backtrack states. Together with effects of force and Gre factors, we demonstrate that recovery from deep backtracks is governed by intrinsic RNA cleavage rather than diffusional Brownian dynamics. We introduce a consensus mechanistic model that unifies our findings with prior models.