4124
Biophysical Journal
Volume 88
June 2005
4124–4136
Braiding DNA: Experiments, Simulations, and Models
G. Charvin,* A. Vologodskii,y D. Bensimon,* and V. Croquette*
*Laboratoire de Physique Statistique, Ecole Normale Supérieure, UMR 8550, Centre National de la Recherche Scientifique, Paris, France;
and yDepartment of Chemistry, New York University, New York, New York
ABSTRACT DNA encounters topological problems in vivo because of its extended double-helical structure. As a consequence, the semiconservative mechanism of DNA replication leads to the formation of DNA braids or catenanes, which have
to be removed for the completion of cell division. To get a better understanding of these structures, we have studied the elastic
behavior of two braided nicked DNA molecules using a magnetic trap apparatus. The experimental data let us identify and
characterize three regimes of braiding: a slightly twisted regime before the formation of the first crossing, followed by genuine
braids which, at large braiding number, buckle to form plectonemes. Two different approaches support and quantify this
characterization of the data. First, Monte Carlo (MC) simulations of braided DNAs yield a full description of the molecules’
behavior and their buckling transition. Second, modeling the braids as a twisted swing provides a good approximation of the
elastic response of the molecules as they are intertwined. Comparisons of the experiments and the MC simulations with this
analytical model allow for a measurement of the diameter of the braids and its dependence upon entropic and electrostatic
repulsive interactions. The MC simulations allow for an estimate of the effective torsional constant of the braids (at a stretching
force F ¼ 2 pN): Cb ; 48 nm (as compared with C ;100 nm for a single unnicked DNA). Finally, at low salt concentrations and
for sufficiently large number of braids, the diameter of the braided molecules is observed to collapse to that of double-stranded
DNA. We suggest that this collapse is due to the partial melting and fraying of the two nicked molecules and the subsequent
right- or left-handed intertwining of the stretched single strands.
INTRODUCTION
The helical structure of double-stranded DNA and the
semiconservative mechanism (Meselson and Stahl, 1958) of
DNA replication pose significant topological constraints
upon DNA (Watson and Crick, 1953): if nothing is done to
separate the newly synthesized strands, they remain
catenated and the process of cell division is stopped.
Furthermore, the progression of the replication complex
induces a torsional constraint in the DNA molecule, which
leads to the formation of supercoils (Wang, 2002; Schvartzman and Stasiak, 2004). A class of enzymes, called
topoisomerases, is in charge of resolving these constraints.
These enzymes are divided into two types: type I and type II
(Wang et al., 1998). Type II topoisomerases use ATP to
induce topological changes in DNA. They transiently cut
a DNA segment (the gate), and pass another segment through
that cut before relegating the gate segment. Using this
mechanism, they can either relax the positive supercoils
generated ahead of the replication fork or unlink the
intertwined daughter DNAs behind. Although twisting
a single DNA molecule offers an interesting in vitro model
for supercoiling, the braiding of two molecules is relevant to
an understanding of the structure of the replication intermediates.
Whereas the behavior of a single DNA molecule under
torsion has been extensively investigated experimentally
(Strick et al., 1996; Cluzel et al., 1996; Leger et al., 1998;
Submitted November 30, 2004, and accepted for publication March 14, 2005.
Address reprint requests to Gilles Charvin, Tel.: 33-1-443-2-3492; E-mail:
[email protected].
Ó 2005 by the Biophysical Society
0006-3495/05/06/4124/13 $2.00
Bryant et al., 2003), numerically (Vologodskii and Cozzarelli, 1996; Vologodskii and Marko, 1997), and theoretically
(Bouchiat and Mézard, 1998; Marko and Siggia, 1995;
Marko, 1997; Moroz and Nelson, 1998a,b), there are but
a few experiments or models (Marko, 1997, 1999) dealing
with the braiding of two molecules. Preliminary experiments
(Strick et al., 1998) have revealed the existence of different
regimes of braiding but did not allow for a comprehensive
interpretation of them. More recent studies have suggested
that DNA braids could form plectonemic structures (Charvin
et al., 2003; Stone et al., 2003), but did not provide direct
evidence of their existence.
In this article, we report an exhaustive experimental and
numerical study of the elastic properties of a braid that
consists of two intertwined nicked double-stranded DNAs.
At high catenation number, the experiment suggests that the
twisted braids undergo a mechanical instability that leads to
the formation of plectonemic structures. Monte Carlo (MC)
simulations confirm this interpretation of the data. These
simulations further allow for a measurement of the torque
acting on the braids and an estimate of their effective
torsional modulus. We show that the braiding of two
molecules is well described by a simple geometric model:
a twisted swing. The diameter of the DNA braid, which
results from both entropic and electrostatic repulsive
interactions, can be deduced by fitting this model to the
experimental data. The variation of the braids’ diameter with
increasing force (which reduces entropic effects) and ionic
strength can thus be investigated. The effective torsional
modulus of the braids deduced from this simple model
doi: 10.1529/biophysj.104.056945
Braiding DNA
4125
agrees with the numerical (MC) estimates. Finally, we suggest that the sharp hysteretic transition observed (Strick et al.,
1998) at low salt results from the melting of the two nicked
DNA molecules, leading to the intertwining of stretched
single strands.
MATERIALS AND METHODS
DNA constructs and micromanipulation setup
We labeled an 11-kb nicked DNA (linking number Lk0 ’ 1100) at its
extremities with biotin and digoxygenin (DIG), as previously described in
Strick et al. (1998). Nicks in the DNA molecules occur randomly during the
various molecular biology steps of the synthesis. Their number as well as
their position is not controlled, although it is not likely that this number
exceeds 10 per molecule, or that the DNA presents large single-strand gaps.
The molecules were bound at one end to a streptavidin-coated, 2.8 mm
magnetic Dynal bead (Dynal Biotech, Lake Success, NY) and at the other to
an anti-DIG-coated glass surface, previously incubated with BSA (Roche,
Nutley, NJ) to reduce nonspecific interactions.
To stretch and braid the molecules, we used small magnets placed above
the sample that can be displaced to set the force pulling on the molecules and
rotated to braid them. Bead tracking and force measurement were performed
on an inverted microscope (Strick et al., 1996). The extension z is measured
by tracking the three-dimensional position of the tethered bead (Strick et al.,
1998), with an error due to Brownian motion of ;10 nm (with 1-s
averaging).
The bead’s transverse fluctuations Ædx2æ allow for a determination of the
stretching force using the equipartition theorem F ¼ kBTz/ Ædx2æ, where kB is
the Boltzmann constant and T the temperature (at 25°C, kBT ¼ 4 pN nm).
Here F was measured with 10% accuracy.
In plots showing the recording of the DNA’s extension, the points
correspond to the raw data and were obtained at 25 Hz. For the curvedisplaying extension versus catenation number n, each point was averaged
over 256 frames. To eliminate microscope drifts, differential tracking with
a second bead glued to the surface was performed.
Geometric model
FIGURE 1 Geometric model for DNA braiding. (a) The case jnj , 0.5,
where simple geometric considerations allow one to derive the extension
z(n) as a function of the angle 2pn, the
length z0 ¼ z(n ¼ 0), and the distance
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2e between the molecules of zðnÞ ¼ z20 4e2 sin2 pjnj. (b) The case jnj .
0.5, where further braiding leads to the formation of a helix with
diameter
Db, so that the extension z(n) follows zðnÞ ¼ z0 cos a ¼
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
z20 ð2e 1 pDb ðjnj 1=2ÞÞ2 . This model remains valid as long as the
molecules are not in close contact, a condition reached when dðs; s9Þ [
~ M9ðs; s9Þk ¼ Db , i.e., when the braid angle a reaches ac ¼ 45° (see
kM
Materials and Methods for details).
nc, the molecule get into close contact, so that the geometric model is no
longer valid. This is described in the next paragraph.
Close contact condition
Two molecules in the braid are not in close contact as long as the distance
d(s, s9) between any two points M(s) and M9(s9) on the molecules’ axes is
larger than D (where s, s9 label the curvilinear coordinates along the molecules’ axes; see Neukirch and van der Heidjen, 2002, for in-depth considerations about the geometry of braids). See Fig. 1 for
0
1
D sinð2psÞ=2
MðsÞ ¼ @ D cosð2psÞ=2 A
ps
and
Simple geometric considerations allow for a description of the braiding of
two DNA molecules. We assume that the DNA molecules of length z0 form
a helix, whose diameter D does not vary with the braiding number n.
Calculation of the extension z versus catenation number n
The regime before contact (jnj , 1/2). Let us define by z0 [ z(n ¼ 0) the
extension of the molecules at n ¼ 0. The variation of the extension z(n) with
the first half-turn (i.e., until the ‘‘ropes’’ are in contact) obeys
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
2
2
zðnÞ ¼ z0 4e sin pjnj:
(1)
The regime of braiding (1/2 , jnj , nc). As the molecules are twisted
beyond n ¼ 1/2, we assume that their braiding results in the formation of
a helical structure of fixed diameter D and braid angle a (Fig. 1), which is
constant all along the molecules, but varies as n changes. A simple
trigonometric argument (see Fig. 1), yields the extension z(n):
zðnÞ ¼ z0 cos a ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
2
z0 ð2e 1 pDðjnj 1=2ÞÞ :
(2)
Notice that the larger n, the larger a. This is in contrast with the structure
of a supercoil, for which the angle of the super-helix is constant when at s .
0.03 (Vologodskii and Cozzarelli, 1996). Beyond a certain number of turns
0
1
D sinð2ps9 1 pÞ=2
M9ðs9Þ ¼ @ D cosð2ps9 1 pÞ=2 A;
ps9
where p ¼ pD/tan a is the braid pitch. The condition d(s, s9) . D ("(s, s9))
remains valid as long as
2
d ðs; s9Þ
1¼
2
D
"
#
2
pðs s9Þ
2
sin pðs s9Þ $ 0;
tan a
which implies tan a # 1, i.e., a # ac ¼ 45°.
Torque G as a function of n
The work Wt performed while twisting the two molecules is equal to the
work done against the force W ¼ F(z0 – z(n)) plus the bending energy of the
braid, Eb ¼ jLbraided ðnÞ=2R2 , where R is the braid effective radius of
curvature and Lbraided(n) is the length of braided DNA . The torque G(n)
stored in the braid is deduced from
GðnÞ ¼
1 @Wt
:
2p @n
(3)
Biophysical Journal 88(6) 4124–4136
4126
Charvin et al.
When n , 1/2, there is no bending energy, so that Eq. 3 yields
2
GðnÞ ¼
@W 2e p sinð2pnÞ
¼
:
@n
zðnÞ
(4)
As the chains are braided for n . 1/2, G(n) becomes
FD tan a
GðnÞ ¼
2
1
j
sin2 aðz0 sin a 1 3pDðn 1=2ÞÞ;
Dz0
(5)
(6)
with sin a ¼ ð2e1pDðn 1=2ÞÞ=z0 .
top, to allow for changes in the chains’ conformation. This was achieved by
adding three potential energy terms, namely Eo, Ed, and Et, to the total energy
of the chain. If ~
t1 ð¼ 2ex̂Þ and ~
t2 are the vector joining the anchoring points,
respectively, at the bottom and the top of the braid, and b the angle between~
t1
and ~
t2 , these terms are
1
2
t2 :ẑÞ
Eo ¼ ko ð~
2
1
2
Ed ¼ kd ðk~
t2 k 2eÞ
2
1
Et ¼ kt b2
2
t2 is orthogonal to the direction of the force, ẑ.
The value Eo ensures that ~
The value Ed prevents the spacing between the molecules from deviating too
much from its average value 2e. The value Et limits the angular fluctuations
of b. Measuring the mean deviation Æbæ allows for an estimation of the
torque in the braid.
Monte Carlo calculations
DNA modeling
The two DNAs were modeled as discrete worm-like chains (WLCs)
constituted by N ¼ 120 rigid segments of fixed length (a ¼ 10 nm), so that
the chain length was L0 ¼ 1.2 mm. If ui is the angle between adjacent segments
i,i 1 1, the bending energy Eb associated to a given chain configuration is
N1
2
Eb ¼ akB T + ui
i¼1
(Frank-Kamenetskii et al., 1985), where kB is the Boltzmann constant, T the
temperature, and a ¼ 2.409 the bending rigidity constant. The value a is set
so that the persistence length of the chain corresponds indeed to five rigid
segments (Frank-Kamenetskii et al., 1985), a discretization which has been
shown to provide an accurate simulation of a DNA molecule (Vologodskii
and Frank-Kamenetskii, 1992). Because of the application of a force F in the
z direction, each segment tends to align along the force, which results in an
additional energy term EF ¼ Fz, where z is the extension of the braid.
Monte Carlo procedure
A Metropolis Monte Carlo algorithm (Metropolis et al., 1953) was used to
generate a set of conformations that satisfy the equilibrium, as described in
Klenin et al. (1991). We randomly applied moves (described in the next
paragraph) to a given braid configuration c (of energy E) to generate a new
trial configuration c9. We calculated the new energy of the braid E9, and
accepted the new configuration with probability P ¼ minf1; expððE9 EÞ=
kB TÞg. This process was then iterated NMC times. A large number of iterations
were required to get a reliable estimate of the braid’s configuration; see below.
Moves
Two types of moves were used to generate new trial configurations: first, we
chose a segment i and changed its angle fi with the z axis to an arbitrary
value (uniformly distributed in cos(fi) 2 [ – 1, 1]), while keeping the
orientation of all other segments fixed. Second, we randomly chose two
vertices i and j in one chain and rotated all the segments between these two
vertices by an angle e around the axis Dij passing through the vertices. The
value of e was randomly chosen in an interval [emax; emax]. The value emax
was set so that the rate of acceptance was 25%. Note that this move did not
change the extension of the chain.
Additional energy term to keep the relative position of the
anchoring points
Although the spacing between the chains was kept fixed at the bottom of the
braids, we had to allow small variations of the spacing and orientation at their
Biophysical Journal 88(6) 4124–4136
Topological constraints
Because of the moves occurring during the Metropolis procedure, the linking
number n between the two chains may change. We selected the conformations
that had the right linking between chains by calculating the Alexander
polynomials for two linked chains (Frank-Kamenetskii and Vologodskii,
1981). Configurations with the wrong linking number were rejected. We also
used Alexander polynomials to evaluate the knotting in the chain formed by
the closing of the two subchains (Frank-Kamenetskii and Vologodskii, 1981)
and rejected all conformations with the wrong topology.
Excluded volume effect and electrostatic interactions
We took into account the excluded volume and electrostatic interactions by
introducing a hardcore diameter DDNA for each segment (Stigter, 1977;
Rybenkov et al., 1993). We discarded configurations whose segments were
intersecting each other.
Measurements of correlations in the Monte Carlo process
To estimate the number NC of Monte Carlo (MC) steps necessary to obtain
decorrelated configurations, we used a method based on the calculation of
the error in the extension mean, as described in depth in Flyvbjerg and
Petersen (1989).
The method relies on the following idea: if the extension z is a random
variable with a mean value z0 and a standard deviation s, these parameters can
be estimated by using the simple formula z0 ¼ Æziæ and s ¼ Æ(zi – z0)2æ 1/2,
where zi is the extension of the configuration of index i. These equalities are
mathematically valid only if two hypotheses are fulfilled: the averaging is
done on a large number of samples N, and these samples are uncorrelated. Of
course, those two conditions are seldom fulfilled. The number of samples N is
always finite, which leads to an error in the estimation of z0N and sN; their
relative accuracy will typically scale like N1/2, the error on z0N being equal
to sN/N1/2. These results are a consequence of the so-called central-limit
theorem.
The correlations between successive zi effects have a more serious impact:
their effects are to reduce the fluctuations of successive zi, leading to an
underestimation of the error on z0N computed by sN/N1/2. A way to evaluate
the correlation is simply to try and check the validity of the central-limit
b
theorem. By using a long series of zi, we build a series of data zN
j
corresponding to the average of zi over Nb consecutive points (Nb ¼ N 3 Q all
integer) with i between Nb 3 j and Nb 3 (j 1 1). We then compute sNb and
1=2
Nb
b
and repeat this process
evaluate the error in z0 as sN
z0 ¼ s =ðN=Nb Þ
while increasing Nb. Two situations occur. If Nb is large enough that the
Braiding DNA
correlation between zi and zi1Nb has disappeared, the central-limit theorem
b
applies, and sN
z0 is independent of Nb. However, if Nb is too small, the
b
correlation reduces the fluctuations and sN
z0 decreases while decreasing Nb.
The transition between these two regimes occurs at a value NC corresponding
to the correlation size of the series. Of course N has to be much larger than NC,
b
otherwise sN
z0 is continuously increasing with Nb and never reaches its
asymptotic value. Typically N needs to be 100 times larger than NC. The ratio
N/NC gives the number of uncorrelated points in the data series.
According to this method, we iteratively computed the error in z0 [ Æzæ
with bins of growing size Nb (corresponding to 10,000 Nb MC steps), as
shown in Fig. 2 a (for the case of n ¼ 6 braids). We obtained an estimate of
the number of correlated MC steps NC (NC ¼ 660,000 for n ¼ 2). NC appears
to vary exponentially with the number of braids n (see Fig. 2 b): NC ¼ A
exp(Bn), where A ¼ 44,000 and B ¼ 0.46, thus indicating that braiding the
chains restricts the ability of the MC algorithm to effectively explore the
phase space and to rapidly reach thermodynamic equilibrium. Consequently,
to ensure a statistically relevant ensemble of independently relaxed
configurations we set the total number of MC steps as N . 100 NC.
EXPERIMENTAL RESULTS
Force versus extension curves for two
unbraided molecules
To study the braiding of two DNA molecules, we have
anchored two nicked 11-kb DNA molecules (crystallographic length L0 ¼ 3.6 mm) on the same superparamagnetic
bead (Strick et al., 1998). This was achieved by incubating
streptavidin-coated magnetic beads (Dynal) with biotin endlabeled DNA at an appropriate DNA/bead ratio. The DNA/
bead construct was injected in an anti-DIG-coated square
capillary tube, to which surface the molecules could bind via
their second extremity labeled with DIG (see Fig. 3).
The DNA/bead ratio was set such that the probability of
anchoring by many molecules was much lower than the
probability of tethering by a single DNA. This ensures that
the few beads anchored by two molecules are much more
frequent than beads anchored by three or more molecules. A
force is applied on the bead by using the magnetic trap setup,
as described by Strick et al. (1998). This setup allows one to
rotate the magnets and thus to braid the two molecules
tethering the bead to the surface. Such doubly anchored
beads can be discriminated easily from beads anchored by
a single DNA molecule. Since the molecules are nicked,
singly anchored beads are insensitive to rotation of the
magnets, whereas multipally anchored beads recoil to the
surface as a result of braiding.
We applied a magnetic force on the bead F by creating
a magnetic field gradient using small magnets in the vicinity
of the glass capillary. This force can be varied by changing
the distance between the magnets and the sample. Once
multipally anchored beads have been identified, we checked
that they were bound by two DNA molecules by measuring
the force-extension response of the system and comparing it
to the WLC model. We hypothesized that there were two
molecules bound in parallel to the bead, so that the force
F1mol felt by each molecule would be half the stretching force
F. Thus, plotting F1mol as a function of extension should
4127
FIGURE 2 Measurements of correlation in the generation of Monte Carlo
configurations. (a) Open circle is the error on mean (d) as a function of binsize Nb (see text for details) obtained at n ¼ 6 (with 2e/L0 ¼ 0.36 and DDNA
¼ 6 nm). Solid
line is the curve obtained by fitting the numerical data to
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dðNb Þ ¼ d0 = 1 1 NC =ð10; 000 Nb Þ. NC yields an estimate of the number
of correlated steps. In the case where n ¼ 6, we obtained NC ¼ 660,000. (b)
Open square is the variation of NC with the number of applied turns n, and
exponential fit as NC ¼ A exp(Bn), with A ¼ 23,000 and B ¼ 0.47. The
higher the braiding number, the larger the number of MC configurations
required to characterize the chain.
agree with the WLC and should output the persistence length
expected for DNA in these conditions. Fig. 4 shows the force
extension curves obtained with one tethering (in this case,
F1mol ¼ F) and two tethering DNA molecules (F1mol ¼ F/2),
respectively. The best fit to the WLC yields j ¼ 46 6 5 nm
and j ¼ 44 6 4 nm, respectively, thus validating that our
bead was doubly anchored in the latter case. A lower
persistent length in the latter case would have been found in
the case of three or more tethering molecules.
Notice that, in principle, the distance between the
anchoring points on the beads and the surface need not be
FIGURE 3 Sketch of the experimental setup. Two DNA molecules are
multipally tagged at their extremities with biotin and digoxygenin (DIG), so
that they can be attached to a streptavidin-coated superparamagnetic bead
and to an anti-DIG-coated glass surface. Translation and rotation of small
magnets close to the DNA molecules by a few-mm change of the applied
force and the number of braids, respectively. The DNA molecule’s extension
is measured by tracking the position of the bead using an inverted
microscope (see Materials and Methods for details).
Biophysical Journal 88(6) 4124–4136
4128
the same, and indeed, constitutes the large majority of sets of
molecules we found in the capillary after injection of the
DNA-bead mix. When the molecules are stretched in such
a trapezoidal conformation, the distance of the bead to the
surface z(F) will be smaller than one single molecule’s
extension at one-half the force L(F/2). Such beads are not
considered in the following, where we focus our attention on
those constructs for which z(F) L(F/2).
Extension z (n) versus catenation number, n
By rotating the magnets above the sample, we are able to braid
the molecules by n turns and to monitor the consecutive
variations in their extension z(n) (see Fig. 5). Three different
regimes can be discerned (see dashed lines in Fig. 5):
1. jnj # 1/2. The first half-turn leads to a big decrease in the
DNA’s extension, which corresponds to the formation of
a crossing between the two DNA. This large decrease is
due to the fact that the distance between the molecules’
anchoring points is comparable to the length of the
molecules. Previous experiments on braided DNA did
not report such an effect because the DNA used was
longer (48 kb, Strick et al., 1998; and 41 kb, Stone et al.,
2003). From the variation of extension between n ¼ 0 and
n ¼ 0.5, we can calculate the spacing between the two
molecules (see Materials and Methods). Note that this
spacing is limited by the size of the bead, i.e., 2.8 mm.
FIGURE 4 Force versus extension curve for one (left plot) and two (right
plot) DNA molecules tethered to the same bead. The value F1mol is the
actual force felt by one molecule, i.e., F1mol ¼ F in case of one tethering
molecule, and F1mol ¼ F/2 in case of two molecules, where F is the
stretching force. Open squares are the experimental points obtained with
a single DNA molecule. Error bars indicate statistical error. The best fit
(solid line) to the worm-like chain (WLC) model yields a persistence length
j ¼ 46 6 5 nm. Open diamonds are the experimental points obtained on
different sets of two DNA molecules; the fit (solid line) using the WLC
model yields j ¼ 44 6 4 nm. Shaded crosses are the results obtained using
a Monte Carlo simulation of the stretching of one (left plot) or two (right
plot) unbraided molecules of persistence length 46 nm, according to the
procedure described in Materials and Methods. Statistical errors of
simulated curves are smaller than the size of the symbols. Small deviations
from the WLC model at forces .2 pN are due to chain discretization
(Vologodskii, 1994).
Biophysical Journal 88(6) 4124–4136
Charvin et al.
2. 1/2 , jnj , nc. Further rotation of the magnets leads to
a smaller decrease in the system’s extension per added
turn. This regime corresponds to the formation of the
braids. As we shall see below, the extension of the molecules in this domain is well described by the simple
geometric model of a twisted swing.
3. jnj . nc. At a critical number of turns nc, the molecules
get in close contact, so that the torque rapidly builds up
and further rotation of the magnets leads to a buckling of
the braids (Marko, 1997). That transition to a different
regime of braiding is manifested by a change in the slope
of the extension versus n plot. That transition is characterized by an increase in the fluctuations in extension,
which is also observed in MC simulations of braided
molecules.
Transition to plectonemes of braids and
longitudinal fluctuations
To get a better understanding of the n . nc regime, we have
monitored the longitudinal fluctuations of the system as
a function of the catenation density, Ca ¼ n/Lk0, defined in
the same manner as the supercoiling density (s ¼ DLk/Lk0).
This definition allows for a comparison between molecules
of different size. Fig. 6 a) displays the extension measured
for one set of molecules as a function of Ca . 0 (shaded
diamond, right-handed braids) and Ca , 0 (open diamond,
left-handed braids). Interestingly, we observe a small but
consistent difference in the molecules’ extension in the
Ca . 0 versus the Ca , 0 regime, which suggests that righthanded braids are more stable than braids of opposite
handedness: they can be twisted to a larger extent before they
buckle. This phenomenon, which is observed only at high
salt, could be due to the fact that the wrapping of two DNA
molecules in a braid with the same chirality as DNA (righthanded) can lead to a more compact structure, thus decreasing the torque stored in the braid.
Fig. 6 b shows the averaged squared fluctuations in
extension Ædz2æ normalized by the molecule’s radius of
gyration: R2g ¼ j L0 =3. Notice the large increase in noise
when Ca . 0.05, which is correlated to a change in the slope
of the extension versus rotation curve. This behavior indicates
that the system’s stiffness decreases at the buckling transition,
consistent with the formation of plectonemes of braids
(similar behavior is observed when stretched twisted DNA
buckles to form supercoils; see Charvin et al., 2004).
MONTE CARLO SIMULATIONS
Force versus extension curves for two
unbraided chains
To simulate the experimental data, we have performed
Monte Carlo simulations (Vologodskii and Marko, 1997) of
the braiding of two discrete WLCs of persistence length j ¼
Braiding DNA
4129
FIGURE 5 Extension versus braiding for
two DNA molecules and sketches of the
expected geometry of the braid. The open
diamonds are experimental points, obtained at
F ¼ 2 pN in 100 mM PB. We distinguish three
regimes of braiding, based on the geometrical
model described in the text: (I) the regime jnj ,
0.5, characterized by a sharp variation with n of
the extension, before the crossing of the two
molecules; (II) the regime jnj . 0.5, where the
molecules are intertwined; and (III) the regime
jnj . nc, where the braids buckle and form
plectonemes (supercoils of braids). Error bars
indicate the statistical errors. A fit of the
experimental data to the geometric model (in
regimes I and II) yield the intermolecular
spacing of 2e ¼ 1.28 6 0.02 mm and braids’
diameter of Db ¼ 8.2 6 0.2 nm (shaded line).
The line in regime III is a linear fit to the data in
the plectonemic regime. The schematics represent the expected geometry of the braid in the
different regimes.
46 nm (a faithful description of the elastic behavior of DNA;
see Vologodskii and Frank-Kamenetskii, 1992; Vologodskii,
1994). Each molecule was modeled as a L0 ¼ 1.2 mm chain
consisting of N ¼ 120 rigid segments of 10-nm length each.
This discretization of the molecule (five segments per
persistence length) provides a good enough approximation
of its entropic behavior at low forces F , 10 pN. The
conformations of DNA molecules were generated using
a Monte Carlo Metropolis algorithm (Metropolis et al.,
1953), as described in Materials and Methods. The
configurational space was randomly sampled by appropriate
moves of the chain. The energy of the sampled configurations consisted of two terms: First, a bending energy
proportional to the square of the angle (ui) between adjacent
segments (i, i 1 1) and calculated by summing over i ¼ 1,
. . . , N; and second, a potential energy term resulting from the
~), which tends to align all the
application of the force (F
segments in its direction (defined as the z axis). A new
configuration was accepted with probability P ¼ 1 if it
lowered the energy of the chain and rejected with an appropriate probability P , 1 (see Materials and Methods) if it
increased it. To check our algorithm we verified that the
simulation of the pulling of two unlinked chains reproduced
the expected elastic behavior of two WLC in parallel. Shaded
symbols in Fig. 4 shows a force-extension curve obtained
using a numerical simulation with one chain or two unlinked
chains. The excellent agreement between these numerical
data and the WLC predictions demonstrates the accuracy of
our algorithm as a model of the elastic behavior of two DNA
molecules.
When the chains are braided, the number of links n
between the two chains must be kept constant during the
simulation. However, as the moves used to generate new
configurations may modify n, we used Alexander polynomials to assess the linking number and reject moves that
altered it (Vologodskii et al., 1975). We applied the same
method to reject knotted chains, although it was not necessary to perform this check at every move.
A reliable sampling of configuration space requires the
generation of a large number of uncorrelated configurations
(the method used to assess the MC correlation length was
described in Materials and Methods). We typically generated
between N ¼ 108 and 4.109 (mostly correlated) configurations, which takes a lot of computer time, especially at
large n (see Fig. 2 in Materials and Methods). As a consequence, we decided to focus on a comparison between
simulation and experimental data for a given separation 2e
between the molecules and DNA’s effective diameter DDNA,
set by the ionic conditions of the experiment (Stigter, 1977).
Extension z (n) versus catenation number n
The extension versus n curves for braided chains greatly
depends on the spacing between the chains. To simulate the
experimental data, we set the spacing 2e between the chains
such that the ratio 2e =L0 ’ 0:36 was the same in the simulation as in the particular experiment shown in Fig. 5. The
electrostatic repulsion between the DNA molecules sets its
effective diameter, which in a solution containing 100 mM of
monovalent salts has been calculated and measured to be
DDNA ¼ 6 nm (Stigter, 1977; Rybenkov et al., 1993, 1997).
To take that repulsion into account we have treated the
chain’s segments as hardcore cylinders of a diameter DDNA.
As shown in Fig. 7 a, a comparison between the simulated
and the experimental data displays a good agreement over
the whole range of catenation densities Ca investigated.
Notice that the extension is sensitive to the value of DDNA:
simulations with DDNA ¼ 5 nm (corresponding to 200 mM
monovalent salt) or 4.2 nm (in 100 mM PB 1 5 mM Mg21)
do not provide as good a fit to the data. However, for Ca .
Biophysical Journal 88(6) 4124–4136
4130
FIGURE 6 Experimental and numerical characterization of the braids’
plectonemic regime. (a) Normalized extension (z/L0 where L0 ¼ 3.56 mm is
the molecules’ length) versus catenation density Ca ¼ jnj/Lk0 for n . 0
(shaded diamonds) and n , 0 (open diamonds; F ¼ 2 pN in 100 mM PB 1 5
mM MgCl2). Error bars indicate the statistical error. Notice the slight
difference between left-handed (L; negative, n , 0) and right-handed (R;
positive, n . 0) braids at high catenation density. This difference reflects the
right-handed chirality of the DNA molecules, which can be positively
braided to a larger extent before buckling than negative (left-handed) braids.
Open circles are the numerical simulations of the braiding of two WLC
polymers of extension L0 ¼ 1.2 mm, intermolecular distance 2e/L0 ¼ 0.36,
and effective diameter DDNA ¼ 4.2 nm, determined from the ionic
conditions of the experiments (Rybenkov et al., 1997). (Dashed lines mark
the differences in slope between the two regimes of braiding.) (b) Averaged
normalized fluctuations in the braid’s extension Ædz2æ/(jL0/3) as a function of
Ca for the experimental (shaded diamonds, n . 0) and the simulated (open
circles) runs shown in a. The fluctuations increase for Ca . 0.05, i.e., the
system becomes more flexible as expected from the formation of supercoils
of braids. The dashed lines mark the differences between the two regimes of
braiding. Notice that the error bars in the estimate of the fluctuations in
extension (from MC simulations) are larger than in the estimate of the mean
extension. The dashed lines in both a and b cross at the same value of Ca,
which indicates that the transition to supercoils of braids is associated with
both a structural change (formation of plectonemes) and a decrease in
stiffness (increase in fluctuations).
0.04 the simulation with a smaller effective radius (DDNA ¼
5 nm) seems to fit the experimental data slightly better. It
should be noted that the concept of the DNA effective
diameter assumes equal probability of all mutual orientation
interacting segments (Vologodskii and Cozzarelli, 1995).
The assumption is not valid for tight DNA braids, and
therefore accounting for the electrostatic repulsion in terms
of DNA effective diameter is less accurate in this case. It is
Biophysical Journal 88(6) 4124–4136
Charvin et al.
also possible that the increased torque in the braids at high
catenation number forces the molecules in closer contact,
reducing the value of DDNA estimated from pure electrostatic
repulsion. This is consistent with the observation, at high
catenation number, of slightly different extensions for rightor left-handed braided DNA. Notice finally the transition in
the simulations to plectonemes of braids at Cac ; 0.04–0.05.
Good agreement between the simulations and the experimental data is also obtained for different ionic conditions
(100 mM PB 1 5 mM Mg1 1 for which DDNA ¼ 4.2 nm;
Vologodskii and Marko, 1997); see Fig. 6 a. Moreover, the
fluctuations in extension for both the experimental and simulated data are similar and exhibit, as expected (see Experimental Results), a larger increase in their amplitude in the
regime of plectonemes of braids. The slightly larger amplitude
of fluctuations in the experiment is probably due to various
external sources of noise, absent from the simulations.
Our simulations therefore provide an excellent description
of the braiding of two DNA molecules. However, whereas in
the experiments we have no control on the distance between
the two braided molecules, in the simulations this is an
adjustable parameter. We can therefore simulate the intertwining of the daughter strands during replication as they are
joined at the replication forks, much more easily than we can
reproduce that configuration experimentally (that would
require a special construct which we have not attempted).
The results of that simulation are shown in Fig. 7 b. At low
catenation number one notices that the braids are localized
near the anchoring points, a configuration that maximizes the
entropy in the rest of the chain (Marko, 1997). This
localization of the topological constraint seems to be a general
feature of random chains also reported for knotted or
catenated plasmids. It has been invoked to explain the
amazing property of the type II topoisomerases, which,
despite acting locally on DNA crossings, are capable of
reducing their topological entanglement (a global property of
the chain(s)) (Marko, 1997; Vologodskii et al., 2001). In the
present instance our simulations suggest that the links between the replicated daughter strands (and therefore probably
the topoisomerases) will be localized near the replication fork.
At high catenation number, the intertwined daughter
strands undergo a buckling transition to plectonemes of
braids. That transition is observed at Cac, 0 ; 0.075, which is
higher than the value Cac ; 0.05 estimated for molecules
anchored a relative distance 2e/L0 ¼ 0.36 from each other.
However, in that case, the effective length of braided DNA is
;36% smaller. Normalizing the number of braids by the
length of the DNA actually being braided yields a value of
Ca9c ; 0.078, close to the value estimated for intertwined
daughter strands.
Force versus Ca phase diagram
In bacteria, the catenation density Ca during the replication
process is difficult to evaluate, since it results from a balance
Braiding DNA
4131
different forces and braiding densities. Based on the analysis
of the braids’ structures, we plot in Fig. 8 the phase diagram
for buckled and unbuckled DNA catenanes. This plot reveals
that the critical force for buckling increases with the braiding
number Ca faster than exponentially.
The supercoiling induced tension in a typical plasmid with
s ¼ 0.06 is F 0.6 pN (Charvin et al., 2004). In such
a plasmid one can estimate from Fig. 8 that buckling of the
replicated daughter strands will occur at a catenation density
of Ca ; 0.065. Previous studies of replication intermediates
(Adams et al., 1992) suggest that their catenation density is
Ca ; 0.01, meaning that plectonemes of catenanes are either
not generated during replication, or are transiently localized
near the replication fork and relaxed by the action of topoisomerases.
Numerical estimation of the braid’s
torsional modulus
FIGURE 7 Relative extension z/L0 versus catenation density Ca in
numerical simulation of the braiding of two polymer chains and comparison
with the experimental data in 100 mM PB (same molecule as in Fig. 8 but in
different ionic conditions). (a) Normalized extension as a function of Ca
obtained in an MC simulation of two chains of extension L0 ¼ 1.2 mm,
intermolecular distance 2e/L0 ¼ 0.36 at F ¼ 2 pN using an effective DNA
diameter DDNA ¼ 6 nm (blue circles, corresponding to the experimental
ionic conditions), DDNA ¼ 5 nm (red circles), and DDNA ¼ 4.2 nm (green
circles). The open diamonds are experimental data. The solid line is a fit of
the numerical data to the geometric model (described in the text) with
a braid’s diameter Db ¼ 8.8 nm . DDNA ¼ 6 nm due to entropic repulsion.
The errors bars in the estimates of the simulated extension are smaller than
the size of the symbols. (b) Numerical results obtained using 2e/L0 ¼ 0.02
(open circles). The solid line is a fit of the numerical data to the geometric
model with Db ¼ 7.4 nm. The schematics display some typical braid
configurations obtained at various Ca values indicated below each structure.
Note that all these simulations were done with Ca . 0 (the braid helix is
right-handed), but identical results were obtained with Ca , 0, since we do
not take DNA chirality in the simulation into account.
between the rate of progression of the replication complex
and the rate of supercoils and pre-catenanes removal by Type
II topoisomerases. Recently, it has been shown that the
bacterial type II topoisomerase, Topo IV, efficiently unlinks
right-handed braided DNA molecules only when they form
plectonemes (Charvin et al., 2003; Stone et al., 2003). It
is thus interesting to get an estimation of the range of forces
and catenation density, Ca, which lead to the formation of
supercoils of braids. To that purpose, we have performed
numerical simulations of braided daughter strands (e ¼ 0) at
The determination of the value of the torsional modulus of
double-stranded DNA is still a subject of controversy, with
values ranging from C ¼ 50 nm kBT and C ¼ 120 nm kBT
depending on ionic conditions (Neukirch, 2004) and the
method used (Charvin et al., 2004). A related question is to
determine the torsional persistence length of a DNA braid
composed of two nicked double-stranded DNAs. In contrast
with the torsional modulus of DNA, which depends upon its
intrinsic material properties, the effective torsional modulus
of braided chains is of entropic origin and can thus be
deduced from MC simulations. Due to the large torsional
stiffness of the magnetic trap it was unfortunately not possible to compare these numerical results to a measurement of
the torsional modulus of DNA braids.
FIGURE 8 Force versus Ca (Ca . 0) phase diagram for the buckling of
braids (with 2e/L0 1). The MC simulations of braids at various forces and
catenation densities (Ca) allow us to determine the limit between buckled
and unbuckled states. The black points display the position of the buckling
transition at a given Ca. Error bars represent statistical error. The thick
shaded line indicates the position of the buckling line. Typical conformations in each phase (buckled or unbuckled) are shown. The dimensionless
unit used on the y axis involved the persistence length j of a single DNA
molecule.
Biophysical Journal 88(6) 4124–4136
4132
Let t 1 and t 2 be the vectors joining the anchoring points
of the two chains at the bottom and the top of the braid,
respectively (see Fig. 9 a). To compute the torque in the
braids from the MC simulations, we do not impose t 1 ¼ t 2,
but let the angle b(p) between these two vectors vary
slightly under the torque acting on the braids. To that purpose we introduce, in the energetics of the chains, an elastic
energy term of Et ¼ ktb2/2 (see Materials and Methods), with
kt ¼ 2000 kBT/rad2 for n , 1/2 and kt ¼ 200 kBT/rad2 for n .
1/2. From the MC simulations we can obtain an estimate of
the average torque in a braided chain as ÆGæ ¼ kt Æbæ, and
from the variation of this torque with n, we can derive the
torsional modulus of the braids as Cb ¼ (L0/2p)@ÆGæ/@n.
Fig. 9, b and c, displays the values of ÆGæ obtained in the
two regimes n , 1/2 and n . 1/2 for the simulation runs
shown in Fig. 7. When n , 1/2, the large spacing between
anchoring points (2e/L0 ¼ 0.36) introduces a large increase
in the torque (solid points), which is maximal at ;n ¼ 1/4
(;25 kBT/rad) and then decreases back to ;2 kBT/rad at
n ¼ 1/2. This sharp variation in the torque is absent when
the molecules are anchored at the same point (or very close
to each other, data not shown).
When n . 1/2, the torque rises almost linearly with n,
whether the spacing between anchoring points is large (2e/L0
¼ 0.36, solid points) or very small (2e/L0 ¼ 0.02, shaded
points). Interestingly, the torque does not seem to saturate
past the buckling threshold, although more data are needed
in this regime, which is hard to simulate. A linear fit to the
numerical results allows us to deduce the braids’ torsional
modulus, Cb/kBT ¼ 63 6 6 nm at a large intermolecular
distance (2e/L0 ¼ 0.36), and Cb/kBT ¼ 48 6 1 nm for chains
anchored at the same point (2e/L0 1). Surprisingly, Cb
Charvin et al.
does not depend much on the spacing between the
molecules. In fact, correcting the value of Cb estimated at
2e/L0 ¼ 0.36 by the length of DNA effectively braided
(;0.64 L0) yields a value C9b ; 41 nm in even better
agreement with the value at 2e/L0 1.
A GEOMETRICAL MODEL FOR DNA BRAIDING
Extension z (n) versus catenation number n
To provide a phenomenological understanding of the
behavior of braided DNA, we have proposed a model that
describes the braiding of the molecules anchored a distance
2e from each other in analogy with the intertwining by n
turns of the two ropes of a swing (see Fig. 1). This model is
a refinement of a model previously proposed for the braiding
of two molecules anchored at the same point (2e ¼ 0) (Strick
et al., 1998). The diameter Db of the braid need not be equal
to the crystallographic diameter of DNA, since entropic
fluctuations and electrostatic repulsion between the two
molecules will tend to increase the intermolecular distance
between the two molecules. The elastic behavior of such
a twisted swing is characterized by the two regimes n , 1/2
and n . 1/2, as described in Materials and Methods (see also
Fig. 1).
When n , 1/2, there is no contact between molecules.
There is an excellent agreement between the braid extension
predicted by Eq. 1 (see Materials and Methods) and our data,
which enables us to estimate the distance 2e between the
molecules’ anchoring points. For the particular data shown in
Fig. 5, e ¼ 0.64 6 0.02 mm. The value z0 is deduced directly
from the value z(n ¼ 0) of the plot.
FIGURE 9 Determination of the torque in the braid. (a) Schematics: the torque is estimated by calculating the deviation angle b between the two vectors t1
and t2 between the anchoring points (see text for details). (b) Regime n , 1/2. The points display the values estimated from the simulations, whereas the solid
line is the torque expected from the geometric model (when 2e/L0 ¼ 0.36, with no fit parameter). (c) The regime n . 1/2. Results from numerical simulations at
2e/L0 (solid squares) and 2e/L0 ¼ 0.02 (shaded squares) and comparison with the estimates from the geometric model (solid and shaded curves, respectively).
Clearly the model fails to predict the variation of G with n, in particular when 2e/L0 1. Errors bars indicate statistical errors. (The solid and shaded dashed
lines are linear fits to the numerical data from which we compute the torsional modulus of the braids, Cb, 63 6 6 nm kBT and 48 6 1 nm kBT, respectively.)
Biophysical Journal 88(6) 4124–4136
Braiding DNA
4133
As the molecules are further braided they form a helical
structure of fixed diameter Db and varying braid angle a. The
very good fit between the extension predicted by this
geometric model (see Eq. 2 in Materials and Methods) and
the data allows us to extract the diameter of the DNA braid
Db (for the data in Fig. 5, Db ¼ 8.2 6 0.2 nm). This value is
larger than the crystallographic diameter of DNA (2 nm), but
also larger than the effective diameter of DNA, DDNA ¼ 6
nm, which was used in the numerical simulations to fit the
experimental data (see above). The reason for this discrepancy is that in the simple model considered here the larger
diameter of the braids results not only from electrostatic
repulsion but also from entropic contributions (which are
naturally taken care of in the MC simulations). As we shall
see below, the geometric model allows us to extract the
respective contributions of these interactions at various
forces and ionic strengths.
Eq. 2 remains valid as long as the molecules are not in
close contact, i.e., when a # ac ¼ 45° (see Materials and
Methods for a proof of that point). Beyond that value, the
torque in the braided molecules leads to their mechanical
buckling, i.e., to the formation of the plectonemes of braids
previously described (see Experimental Results, and Monte
Carlo Simulations, above).
tive diameter of DNA in these ionic conditions—which is
DDNA ¼ 6 nm.
Estimating the torque in braided chains from the
geometric model
The numerical estimates of the torque in braided chains can
be understood within the framework of the simple geometric
model previously described. Under the assumptions of this
model, the work Wt performed while twisting the two DNA
molecules by n turns is balanced by two terms: first, the work
W done against the force pulling on the molecules and
second, the bending energy of the braided DNA molecules.
From the calculation of that energy (detailed in Materials
and Methods), one can deduce the torque stored in the braid
G(n) ¼ 1/(2p)@Wt/@n.
When n , 1/2, there is no contact between molecules and
no bending. The torque is due purely to the work against the
Braid diameter versus force and salt
We have experimentally investigated the influence of
entropic and electrostatic repulsion interactions on the braid
diameter Db. We can modulate the strength of the entropic
repulsion between the molecules by varying the stretching
force: the larger the force, the smaller the fluctuations of the
molecules; the smaller their entropic repulsion, the smaller
Db is. By changing the ionic conditions, we similarly
modulate the electrostatic repulsion range (Debye length)
between the molecules. The lower the ionic strength, the
larger the Debye length, and the larger the Db.
Fig. 10, a and b, display the variation of extension with
the number of braids n obtained in different phosphate
buffers and at different forces. Fitting the various curves to
the geometric model previously introduced allow us to
retrieve the braid diameter as a function of force and ionic
strength (see Fig. 10 c). Increasing the tension reduces the
entropic fluctuations and decreases the braid diameter. At
any given salt conditions, the dependence of Db on the
force is well fit (see Fig. 10 c) by the prediction of Marko
(1997), as
3=4
Db ¼ Db;N 1 AF
;
where Db, N is the braid diameter at very large forces, in
which only electrostatic repulsion should play a role. As
shown in Fig. 10, Db, N increases linearly with the Debye
length lD (Zhang et al., 2001). Note that in 100 mM PB,
Db, N ¼ 5.6 6 0.3 nm, close to the value of the effec-
FIGURE 10 Dependence of the braid diameter Db on force and ionic
conditions. (a) Extension versus n curves in 10 mM phosphate buffer (PB) at
different stretching forces and fits using the geometrical model (solid lines);
red circle, 4 pN; green diamond, 2 pN; light-blue circle, 1 pN; and dark-blue
circle, 0.5 pN. (b) Extension versus n curves at F ¼ 1 pN and fits using the
geometrical model (solid lines); red circle, 1 mM PB; light-blue circle, 10
mM PB; and dark-blue circle, 100 mM PB. (c) Evolution with the force of
the fitted diameter of the braids Db at different ionic concentrations: red
circle, 1 mM PB; blue circle, 10 mM PB; green circle, 100 mM PB; and
magenta circle, 100 mM PB 1 5 mM MgCl2. Error bars indicate the error
(determined by bootstrap test) on the diameter fitted using the geometrical
model. Following Marko (1997), Db was fitted by Db ¼ AF3/4 1 Db,N. (d)
Evolution of Db,N (circle) with the Debye length of lD, calculated by Zhang
et al. (2001) at the corresponding ionic concentration. Error bars indicate the
error on Db,N, which has been fitted using least-square methods. A linear fit
yields Db,N ¼ 2.5 lD 1 3.6 nm.
Biophysical Journal 88(6) 4124–4136
4134
force and can be computed from the variation of the braid’s
extension between n ¼ 0 and n ¼ 1/2. Fig. 9 b shows the
value of G(n) calculated for the parameters used in the
simulations, see Fig. 7 a, (2e/L0 ¼ 0.36). The agreement
between the geometric model and the simulation is very
good. However, note that, because of the chains’ fluctuations
(that are neglected in the geometric model), there is a slight
discrepancy near n ¼ 1/2; the chains come into contact
before n ¼ 1/2, thus increasing the torque.
When n . 1/2, the computed torque G(n) has a marked
nonlinear dependence with n, which differs significantly from
the results of the MC simulations (see Fig. 9 c). A possible
reason for that discrepancy might be the finite persistence
length of the braided chains, which would restrict their
bending when the braid pitch is of similar size. The braid
angle a in the model varies continuously until its buckling at
a ¼ 45°, but the value of a deduced from the MC simulations
(with persistence length j ¼ 50 nm) varies little with n.
Although this effect does not alter the variation of z with n, it
does affect the estimation of the energetics of bending in the
braiding regime and therefore the calculation of Cb.
THE COLLAPSE TRANSITION OF BRAIDS AT
LOW SALT
The braiding of nicked DNA molecules in low salt
conditions (#10 mM PB) exhibits a surprising transition to
a tight braided structure (Strick et al., 1998). Fig. 11 a shows
that, in low salt, once the molecules have adopted the braided
plectonemic structure described previously they may undergo a sharp hysteretic transition to a state with a larger
extent and a much smaller decrease in extension upon
braiding. In this regime, the extension varies much less per
added turn (dz/dn ¼ 4 nm/turns) than in the regime of (nonplectonemic) braids at small n (dz/dn 32 nm/turns). The
decrease in extension being proportional to the diameter of
the braided structure, this result suggests that the intertwined
Charvin et al.
molecules in the collapsed regime wound around each other
in a structure much tighter than the regular braid between
two dsDNAs. Since in the conditions of this experiment
(F ¼ 2 pN, 1 mM PB), the braid diameter Db ¼ 25 nm, see
Fig. 10 c, we estimate by simple proportionality that the distance between the molecules in the collapse regime is Dcoll ¼
Db 3 4/32 3.1 nm, which is close to the crystallographic
diameter of DNA.
This collapse transition is observed for both right-handed
(n . 0) and left-handed (n , 0) braids, although it appears to
happen for a slightly smaller value of jnj for left-handed turns.
Surprisingly, this collapse is observed only in low salt
conditions and not at high salt or in presence of Mg21, where
one might have expected electrostatic repulsion to be weaker
and the molecules to be closer to each other to begin with. The
transition is reversible but highly hysteretic. The hysteresis is
more pronounced for left- than for right-handed braids.
Fig. 11 b is a real-time recording of the transition to the
collapsed state of a braid. The collapse occurred ;2 min after
the molecules had been braided by n ¼ 35 turns. Notice that
the transition is very quick (although slower than our acquisition rate, 25 Hz) and that the molecules can be reversibly
unbraided from the collapse regime by simply rotating the
magnets back to zero. Analysis of the longitudinal fluctuations
at n ¼ 26 (data not shown) before and after the collapse
transition showed that they decreased significantly after the
transition, suggesting that the collapsed state is not only tighter
but also stiffer than a regular braid (or a braid supercoil).
DISCUSSION
The braiding of two DNA molecules can be studied in great
detail using a magnetic trap technique. Qualitatively and
quantitatively the molecules respond to braiding as would
the ropes in a twisted swing. Monte Carlo simulations of two
braided WLC polymers fit the experimental data over the
whole range of observation and confirm the transition to
FIGURE 11 Torque-induced collapse of the braids.
(a) Extension z versus catenation number (n) at 1 mM
PB and F ¼ 2 pN, with increasing (or decreasing) n
shaded (or open) circles. At sufficiently high jnj (in the
plectonemic regime of braids), a hysteretic transition
occurs (with a noticeable jump in extension; see arrows
at n ¼ 34 and n ¼ 35) to a state characterized by
a smaller decrease in extension per turn (Strick et al.,
1998), dz/dn. In that state (i.e., for jnj . 35), dz/dn ’ 4
nm/turn, whereas it is ’ 32 nm/turn at small n. This
observation suggests that the structure of braids in that
state is very compact (with a braid diameter approximately eight times smaller than for small n). (b) Realtime observation of braid collapse. Recording of the
extension z (open squares) and the set value of n
(continuous line) as a function of time. Notice the quick
and spontaneous increase in z at n ¼ 35. Unlinking
the molecules back to n ¼ 0 reproduces the initial
configuration.
Biophysical Journal 88(6) 4124–4136
Braiding DNA
plectonemes of braids at high braiding number. The elastic
behavior of two braided polymers deduced from these
simulations is also well fitted by the model of a twisted
swing. As far as the experimental data is concerned, the
length of each molecule is given by the extension of a WLC
polymer under half the exerted tension, F/2. The diameter of
the braid is set by the repulsive electrostatic and entropic
interactions. The variation of the diameter with force is well
fit by the predictions of Marko (1997), namely it decreases
as F3/4 due to a reduction in entropic fluctuations at high
forces. The electrostatic repulsion increases the DNA’s effective diameter by an amount proportional, as expected, to
the interaction range: the Debye length.
When the persistence length of the chains is comparable to
the pitch of the braids, the numerical results show that the
braid angle a differs from the predictions of our simple
model. It saturates at an angle ac less than the predicted value
of 45° (data not shown). This numerical observation results
from the stiffness of the chains that prevents them from
bending as they wind around each other with an angle
smaller than predicted. This effect due to the finite persistence length of the chains appears not to significantly affect
the extension versus catenation behavior of braided chains,
which may partly explain the good agreement between our
simple geometric model and both MC simulations and
experimental data. However, the difference observed between the torque expected from the geometric model and the
one deduced from numerical simulation shows that the
bending modulus of DNA, which is not taken into account in
the geometric model, is important to understand the torque in
DNA braids.
At large braiding number and in low salt conditions, we
have observed a sudden transition to a structure still affected
by the braiding of the molecules but with a decrease in
extension per added turn corresponding to the molecules
being a distance ;3.1-nm apart. It is surprising that such
a transition is only observed in low salt conditions (1 mM PB
and F . 2 pN or 10 mM PB and F . 4 pN), where
electrostatic repulsive interactions have a much larger range
(lD 6 nm). It is probable that, in these conditions, the
nicked DNA molecules are frayed, which may allow the
single strands under tension to wind around each other in a
tight, right- or left-handed, double-helical structure, as described on Fig. 12. This would explain the sharp transition
observed when increasing the braiding number n as well as
the smooth recovery to the initial structure when going back
to n ¼ 0. Note that the detailed structure of this braid greatly
depends on the number of the nicks and their positions in the
DNA molecule. Repeating the experiment with different sets
of molecules lead to slightly different jumps in extension
upon collapse as well as different braiding thresholds (data
not shown). However, the experiment presented in Fig. 11 is
the typical behavior of the braid observed in these conditions. A further study should investigate this transition
quantitatively, using a well-defined specifically nicked con-
4135
FIGURE 12 Model for the collapse transition observed in low ionic
conditions. At small linking number, nicked double-stranded DNAs wrap
around each other, thus reducing the system’s extension (center). Beyond
a critical linking number, the torque in the braids induces the melting of the
molecules and the extrusion of single strands (right). The remaining (braided
and stretched) single strands wind around each other in a helix that is tighter
than the braid formed by two dsDNAs, thus increasing sharply the system’s
extension.
struct. For example, it would be of great interest to determine
the torque required to fray the molecules. Last, it has not
escaped our notice that a critical test of the structure we
suggest in Fig. 12 would be to try the relaxation of the singlestrand braid by type Ia topoisomerase, which is known to
operate single-strand passages. However, such a study is
well beyond the scope of this article.
The results of the present experiments, simulations, and
modeling shed new light on the structures (braids, plectonemes of braids, collapsed structures) possibly adopted by
the catenated sister chromosomes in replication intermediates.
We acknowledge helpful discussions with J. Marko, J.-F. Allemand, S.
Neukirch, and Y. Zhang and useful comments by the referees of this article.
We also thank G. Lia for DNA constructs.
This work was supported for G.C., D.B., and V.C. by grants from
Association de Recherche sur le Cancer, Centre National de la Recherche
Scientifique, Ecole Normale Supérieure, Universities Paris VII and VI,
European Union (‘‘MolSwitch’’), and for A.V. by the National Institutes of
Health grant No. GM54215.
REFERENCES
Adams, D., E. Shekhtman, E. Zechiedrich, M. Schmid, and N. Cozzarelli.
1992. The role of topoisomerase IV in partitioning bacterial replicons and
the structure of catenated intermediates in DNA replication. Cell. 71:
277–288.
Bouchiat, C., and M. Mézard. 1998. Elasticity theory of a supercoiled DNA
molecules. Phys. Rev. Lett. 80:1556–1559.
Bryant, Z., M. D. Stone, J. Gore, S. B. Smith, N. R. Cozzarelli, and C.
Bustamante. 2003. Structural transitions and elasticity from torque measurements on DNA. Nature. 424:338–341.
Charvin, G., V. Croquette, and D. Bensimon. 2003. Single molecule study
of DNA unlinking by eukaryotic and prokaryotic type II topoisomerases.
Proc. Natl. Acad. Sci. USA. 100:9820–9825.
Biophysical Journal 88(6) 4124–4136
4136
Charvin et al.
Charvin, G., T. Strick, J.-F. Allemand, V. Croquette, and D. Bensimon.
2004. Twisting DNA: single molecules studies. Contemp. Phys. 45:383–
403.
Rybenkov, V., C. Ullsperger, A. Vologodskii, and N. Cozzarelli. 1997.
Simplification of DNA topology below equilibrium values by type II
topoisomerases. Science. 277:690–693.
Cluzel, P., A. Lebrun, C. Heller, R. Lavery, J.-L. Viovy, D. Chatenay, and
F. Caron. 1996. DNA: an extensible molecule. Science. 271:792–794.
Schvartzman, J. B., and A. Stasiak. 2004. A topological view of the
replicon. EMBO J. 5:256–261.
Flyvbjerg, H., and H. G. Petersen. 1989. Error estimates on averages of
correlated data. J. Chem. Phys. 91:461–466.
Stigter, D. 1977. Interactions of highly charged colloidal cylinders with
applications to double-stranded DNA. Biopolymers. 16:1435–1448.
Frank-Kamenetskii, M., A. Lukashin, V. Anshelevich, and A. Vologoskii.
1985. Torsional and bending rigidity of the double helix from data on
small DNA rings. J. Biomed. Struct. Dyn. 2:1005–1012.
Stone, M. D., Z. Bryant, N. J. Crisona, S. B. Smith, A. Vologodskii, C.
Bustamante, and N. R. Cozzarelli. 2003. Chirality sensing by
Escherichia coli topoisomerase IV and the mechanism of type II
topoisomerases. Proc. Natl. Acad. Sci. USA. 100:8654–8659.
Frank-Kamenetskii, M., and A. Vologodskii. 1981. Topological aspects of
the physics of polymers: the theory and its biological applications. Sov.
Phys. Usp. 24:679–696.
Klenin, K. V., A. V. Vologodskii, V. V. Anshelevich, A. M. Dykhne, and
M. D. Frank-Kamenetskii. 1991. Computer simulation of DNA supercoiling. J. Mol. Biol. 217:413–419.
Strick, T., J. Allemand, D. Bensimon, A. Bensimon, and V. Croquette.
1996. The elasticity of a single supercoiled DNA molecule. Science.
271:1835–1837.
Strick, T., J.-F. Allemand, D. Bensimon, and V. Croquette. 1998. The
behavior of supercoiled DNA. Biophys. J. 74:2016–2028.
Leger, J. F., J. Robert, L. Bourdieu, D. Chatenay, and J. F. Marko. 1998.
RecA binding to a single double-stranded DNA molecule: a possible role
of DNA conformational fluctuations. Proc. Natl. Acad. Sci. USA. 95:
12295–12299.
Vologodskii, A. 1994. DNA extension under the action of an external force.
Macromolecules. 27:5623–5625.
Marko, J. 1997. Supercoiled and braided DNA under tension. Phys. Rev. E.
55:1758–1772.
Vologodskii, A., and M. Frank-Kamenetskii. 1992. Modeling supercoiled
DNA. Methods Enzymol. 211:468–472.
Marko, J. 1999. Coupling of intramolecular and intermolecular linkage
complexity of two DNAs. Phys. Rev. E. 59:900–912.
Vologodskii, A., A. Lukashin, and M. Frank-Kamenetskii. 1975.
Topological interactions between polymer chains. Sov. Phys. JETP. 40:
932–936.
Marko, J., and E. Siggia. 1995. Statistical mechanics of supercoiled DNA.
Phys. Rev. E. 52:2912–2938.
Meselson, M., and F. W. Stahl. 1958. The replication of DNA. Cold Spring
Harb. Symp. Quant. Biol. 23:9–12.
Metropolis, N., A. Rosenbluth, M. Rosenbluth, and A. Teller. 1953. Simulated annealing. J. Chem. Phys. 21:1087–1092.
Moroz, J., and P. Nelson. 1998a. Entropic elasticity of twist-storing polymers. Macromolecules. 31:6333–6347.
Moroz, J., and P. Nelson. 1998b. Torsional directed walks, entropic
elasticity and DNA twist stiffness. Proc. Natl. Acad. Sci. USA. 94:
14418–14422.
Neukirch, S. 2004. Extracting DNA-twist rigidity from experimental
supercoiling data. Phys. Rev. Lett. 93:198107-1–198107-2.
Neukirch, S., and G. van der Heidjen. 2002. Geometry and mechanics of
uniform N-plies: from engineering ropes to biological filaments. J. Elast.
69:41–72.
Rybenkov, V., N. Cozzarelli, and A. Vologodskii. 1993. Probability of
DNA knotting and the effective diameter of the DNA double-helix. Proc.
Natl. Acad. Sci. USA. 90:5307–5311.
Biophysical Journal 88(6) 4124–4136
Vologodskii, A., and N. Cozzarelli. 1995. Modeling of long-range electrostatic interactions in DNA. Biopolymers. 35:289–296.
Vologodskii, A., and J. Marko. 1997. Extension of torsionally stressed
DNA by external force. Biophys. J. 73:123–132.
Vologodskii, A. V., and N. R. Cozzarelli. 1996. Effect of supercoiling on
the juxtaposition and relative orientation of DNA sites. Biophys. J. 70:
2548–2556.
Vologodskii, A. V., W. Zhang, V. V. Rybenkov, A. A. Podtelezhnikov, D.
Subramanian, J. D. Griffith, and N. R. Cozzarelli. 2001. Mechanism of
topology simplification by type II DNA-topoisomerases. Proc. Natl.
Acad. Sci. USA. 98:3045–3049.
Wang, J. C. 2002. Cellular roles of DNA topoisomerases: a molecular
perspective. Nat. Rev. Mol. Cell Biol. 3:430–440.
Wang, M., M. Schnitzer, H. Yin, R. Landick, J. Gelles, and S. Block. 1998.
Force and velocity measured for single molecules of RNA polymerase.
Science. 282:902–907.
Watson, J., and F. Crick. 1953. Molecular structure of nucleic acids:
a structure for deoxyribose nucleic acid. Nature. 171:737–738.
Zhang, Y., H. Zhou, and Z.-C. Ou-Yang. 2001. Stretching single-stranded
DNA: interplay of electrostatic, basepairing, and base-stacking interactions. Biophys. J. 81:1133–1143.