2.DNA-DNA Interactions in TIght Supercoils Are Described by A Small Effective Charge Density

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 4

week ending

PRL 105, 158101 (2010) PHYSICAL REVIEW LETTERS 8 OCTOBER 2010

DNA–DNA Interactions in Tight Supercoils Are Described by a Small Effective Charge Density
Christopher Maffeo,1 Robert Schöpflin,2 Hergen Brutzer,3 René Stehr,2 Aleksei Aksimentiev,1,*
Gero Wedemann,2,† and Ralf Seidel3,‡
1
Department of Physics, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA
2
CC Bioinformatics, University of Applied Sciences Stralsund, 18435 Stralsund, Germany
3
Biotechnology Center Dresden, University of Technology Dresden, 01062 Dresden, Germany
(Received 12 May 2010; published 4 October 2010)
DNA-DNA interactions are important for genome compaction and transcription regulation. In studies of
such complex processes, DNA is often modeled as a homogeneously charged cylinder and its electrostatic
interactions are calculated within the framework of the Poisson-Boltzmann equation. Commonly, a charge
adaptation factor is used to address limitations of this theoretical approach. Despite considerable theoretical
and experimental efforts, a rigorous quantitative assessment of this parameter is lacking. Here, we
comprehensively characterized DNA-DNA interactions in the presence of monovalent ions by analyzing
the supercoiling behavior of single DNA molecules held under constant tension. Both a theoretical model
and coarse-grained simulations of this process revealed a surprisingly small effective DNA charge of 40% of
the nominal charge density, which was additionally supported by all-atom molecular dynamics simulations.

DOI: 10.1103/PhysRevLett.105.158101 PACS numbers: 87.14.gk, 82.37.Rs, 87.15.A, 87.15.La

The large linear charge density is a fundamental prop- tweezers experiments [14–16]. An illustration of the ex-
erty of DNA which governs its biological function by periment is shown in Fig. 1(a). A DNA molecule is tethered
influencing DNA folding, packaging [1], pairing [2], and between a glass surface and a 1:0 m magnetic bead (see
interactions with other biological macromolecules [3]. In supplementary material [17]). Nearby magnets allow us to
order to develop meaningful quantitative models describ- stretch and twist the attached molecule. Initially, the mea-
ing such systems and processes, a precise knowledge of the sured DNA end-to-end distance reduces only slightly upon
interaction between two DNA molecules, which is mostly supercoiling. Once a critical supercoil density is reached,
of electrostatic origin, is mandatory. DNA electrostatics is the molecule buckles and the end-to-end distance LDNA
affected by surrounding counterions, which screen the decreases linearly with the number of added turns N
DNA charge on the scale of the Debye length D . The [Fig. 1(b)] as the extrusion of a superhelical structure
counterion cloud is mainly set by the interplay between absorbs the additional turns in the form of writhe [18].
solute-ion electrostatic attraction and entropic repulsion, While the torque within the DNA increases linearly before
for which the Poisson-Boltzmann (PB) equation provides a the supercoiling transition, it remains constant afterwards
mean-field description. For highly charged polymers such [14,19]. The slope dLDNA =dN in the superhelical phase
as DNA, this approach can have considerable limitations depends on the applied force and on the ionic strength of
arising from the reduced structural detail with which the the solution [Fig. 1(c)]. Several models have been devel-
DNA macromolecule is approximated and the assumption oped to describe this dependence theoretically [20,21], but
of a continuous counterion density. To offset these prob- a quantitative prediction of the slopes has not yet been
lems, it is common practice to rescale DNA electrostatic achieved [14,21] (see Fig. S1 [17]).
potentials with a charge adaptation factor. Values between In order to provide an improved description of the super-
70% and 100% of the bare DNA charge density are typi- helical regime, we calculate the energy Esh
tot per added turn
cally used [4–6], but the correct parameterization is de- to form an ideal DNA superhelix with superhelical radius
bated [4]. Often such factors are indirectly obtained from  and helical repeat length h in the absence of fluctu-
electrophoresis experiments [7], which are unrelated to ations [4,5,21] [Fig. 1(a)], with the DNA charge as a free
DNA-DNA interactions [8,9]. Direct experimental studies parameter:
of DNA-DNA interaction are rare and themselves limited,
tot ¼ Epot þ Ebend þ EEstat :
Esh force DNA DNA
(1)
e.g., to condensed DNA phases [10], short DNA
particles [11], or large distances [12,13]. Despite increas- The value Eforce
pot denotes the potential energy change due to
ingly sophisticated experiments, an unambiguous quanti- shortening the DNA end-to-end distance against the
tative assessment of DNA-DNA interactions has not been applied force F, EDNAbend the bending energy of the DNA
achieved yet. within the superhelix, and EDNAEstat the DNA-DNA electro-
Here we address this issue by analyzing the ionic static interaction energy. Per added turn, i.e., per super-
strength-dependent supercoiling response of single DNA helical writhe, the DNA length within the superhelix grows
molecules held under constant tension in magnetic by [5] ½ð2Þ2 þ h2 =h ¼ dLDNA =dN, which equals the

0031-9007=10=105(15)=158101(4) 158101-1 Ó 2010 The American Physical Society


week ending
PRL 105, 158101 (2010) PHYSICAL REVIEW LETTERS 8 OCTOBER 2010

a b nominal charge density. The electrostatic potential along


600
N S one DNA double strand is obtained by integrating point

DNA length (nm)


charge potentials placed at the center line of the opposite
Magnets F
400 double strand. The DNA charge density  entering the
potential is adapted (see Fig. S2 [17]) to account for
h the cylindrical geometry (factor Rod ) and for deviations
200
DNA due to using solutions to the linearized PB equation (factor
2ρ 0 5 10 15 PB ). For an ideal infinite superhelix the electrostatic
Turns
potential is invariant along the DNA contour, and the
c electrostatic interaction energy per added turn is approxi-
70 1.00
70 Slope (nm) 0.55 mately the product of the calculated potential (see Fig. S2
60 CR
0.42
0.32
[17]) and the adapted charge of one of the double strands
60
50
[22], 1=2dLDNA =dN :
40
Slope (nm)

30 60 mM Na+ 1 dLDNA Z erðsÞ=D


50
Estat ¼
EDNA kB TlB 2 ds; (4)
0 1 2 3 4 2 dN s rðsÞ
Force (pN)
40
with  ¼ CR Rod PB and lB ¼ 0:7 nm being the
Bjerrum length in water. rðsÞ denotes the distance between
30
CR= 0.42
a point located at position s of the center line of one double
20
strand and a fixed point located at the center line of the
0 1 2 3 4 other double strand of the superhelix. Numeric integration
Force (pN)
is performed over the whole center line s of the former
FIG. 1 (color). Dependence of DNA supercoiling on force and double strand. This simplified electrostatic potential yields
salt concentration. (a) Experimental setup and superhelix pa- values similar to the full numerical solution of the PB
rameters. (b) DNA supercoiling curves recorded in buffer con- equation (see Fig. S3 [17]).
taining 170 mM Naþ at stretching forces of 0.25, 0.5, 1.0, 2.0, By combining Eqs. (1)–(4) we now can calculate Esh tot ,
3.0, and 4.0 pN (gray, light blue, dark blue, red, green, and dark and by minimization with respect to  and h one obtains
gray lines, respectively) for a 1.9 kbp long DNA molecule. the slope dLDNA =dN and the DNA torque  ¼ Esh tot =2 for
Continuous twisting was carried out at 0.5 Hz. Data were taken the energetically favored superhelix configuration.
at 300 Hz and smoothed to 20 Hz. (c) Slopes after buckling as a We compared the predictions of our model with the
function of force obtained from the supercoiling curves (taken in experimentally obtained slopes of our supercoiling curves
part from previous measurements [16]) for Naþ concentrations
for forces between 0.25 and 4 pN [Fig. 1(b)] in buffers
of 30, 60, 170, and 320 mM (gray, red, black, and blue circles,
respectively) together with the prediction from the theoretical
containing 30–320 mM Naþ . Remarkably, the slopes at all
model (solid lines) calculated for a charge adaptation factor CR applied forces and ion concentrations are accurately
of 0.42. Inset: Slopes for 60 mM Naþ (red circles) together with described [Fig. 1(c)]. However, achieving the agreement
theoretical predictions for different values of CR (lines). required a substantial reduction of the DNA charge by
employing CR ¼ 0:42, independent of the salt concentra-
tion (see Fig. S4 [17]). Our theoretical model accurately
slope of the supercoiling curves when neglecting fluctua- describes the slopes of recently available, independently
tions. Eforce
pot is then given by measured supercoiling curves [14] [Fig. 2(a)]. Since the
pot ¼ FdLDNA =dN:
Eforce (2)
a b 30
70 50 mM
The bending energy per turn can be written as [5]
Torque (pN nm)

100 mM
Slope (nm)

200 mM
 2
60
500 mM 20
dLDNA 1 
¼
50
EDNA pkB T ; (3)
bend
dN 2 2 þ ðh=2Þ2 40
10
30
where p ¼ 50 nm denotes the bending persistence length,
20 0
kB the Boltzmann constant, and T the absolute tempera- 0 1 2 3 4 0 1 2 3 4
Force (pN) Force (pN)
ture. The term within square brackets describes the DNA
curvature. We calculate the electrostatic interaction energy FIG. 2 (color). Comparison of the predictions from the theo-
between the two DNA double strands of the superhelix as retical model with data from Ref. [14]. (a) Slopes after buckling
previously described [22,23]. DNA is approximated as a versus force for different Naþ concentrations. Circles represent
cylindrical, uniformly charged rod with 1.2 nm radius experimental data and solid lines the theoretical prediction for
and linear charge density CR , where CR is the adjust- CR ¼ 0:42. (b) Torque after buckling as calculated from force-
able charge adaptation factor and  ¼ 2e=0:34 nm is the extension data [14]. Colors and symbols are as in (a).

158101-2
week ending
PRL 105, 158101 (2010) PHYSICAL REVIEW LETTERS 8 OCTOBER 2010

force dependence of slope and torque are linked [24], our a 30 mM 320 mM

model also reproduces the torque in the superhelical phase


as obtained from force-extension data [14] [see Figs. 2(b) b c
and S5 [17]]. The obtained CR is found to be rather

Slope (nm)
600 70 0.55
70 CR 0.42

DNA length (nm)


insensitive to other parameters of the model; e.g., changing 50 0.32

Slope (nm)
60
the weakly salt-dependent persistence length by 5 nm alters 400 30
50
CR only by 0:03. 0 1 2 3
Force (pN)
4

40
To exclude the possibility that the small effective charge
200 30
is an artifact due to neglected fluctuations, we carried out CR= 0.42
coarse-grained Monte Carlo simulations (Fig. 3) [25]. The 0 5 10 15
20
0 1 2 3 4
DNA is modeled as a chain of small cylindrical segments, Turns Force (pN)

with equivalent terms describing the potential, bending, d e

Torque plateau (pN nm)


and electrostatic energies as in the theoretical model [17]. 30 30

Torque (pN nm)


Additional terms account for the twisting and stretching
energy [6]. By performing simulations with different num- 20 20

bers of added turns, we obtained supercoiling curves very


10 10
similar to those observed experimentally [Fig. 3(b)].
By applying CR ¼ 0:42, the slopes from simulation, the-
0 0
ory, and experiments are in excellent agreement over the 0 5 10 15 0 1 2 3 4
Turns Force (pN)
entire range of applied forces and ionic strengths [Fig. 3(c)].
This provides strong, independent support for a small FIG. 3 (color). Coarse-grained Monte Carlo simulations of
effective DNA charge. Comparing torque values from the DNA supercoiling. (a) Snapshots of the formed plectoneme at
simulations and theoretical model, we find good overall 1 pN and 8 turns in the presence of 30 and 320 mM monovalent
agreement, except the values from simulations appear ions. (b) Simulated DNA supercoiling curves for 170 mM Naþ at
globally 1.5 pN nm higher [Fig. 3(e)]. We attribute this stretching forces of 0.25, 0.5, 1.0, 2.0, 3.0, and 4.0 pN for CR ¼
deviation to DNA fluctuations which are neglected in the 0:42 [colors are as in Fig. 1(b)]. Simulations reproduce the
model (see Figs. S6–S8 [17]). experimentally observed abrupt buckling at the onset of the
plectonemic phase, which is accompanied by a torque overshoot
In order to obtain a microscopic verification for CR , all-
[see (d)]. Buckling is followed by a linear DNA length decrease
atom molecular dynamics simulations [26] were employed with added turns at constant torque. (c) Slopes after buckling
to obtain the force between parallel DNA molecules at obtained from the simulated curves for Naþ concentrations of
different salt concentrations [17]. We obtained good agree- 30, 60, 170, and 320 mM [circles, colors as in Fig. 1(b)] and
ment with the force calculated according to Eq. (4) using corresponding predictions from the theoretical model for CR of
CR ¼ 0:42 and very poor agreement with CR ¼ 1:0 0.42 (solid lines). Inset: Slopes from simulated curves at 60 mM
(Fig. 4). To test whether the value found for CR is specific Naþ for CR ¼ 0:42 (red dots) as well as for CR ¼ 0:32 and
to DNA-DNA interactions or is a universal constant for CR ¼ 0:55 (gray dots with dashed lines). The prediction for
DNA electrostatics, we obtained the ion distributions CR ¼ 0:42 is shown as a solid red line. (d) Torque during DNA
supercoiling for the curves shown in (b). (e) Torque after buck-
around isolated double-stranded DNA molecules. The ra-
ling as obtained from the simulations (filled circles) and after
dial ion distribution extended further from the DNA than subtraction of 1.5 pN nm (open squares) together with the
expected from PB theory with CR ¼ 0:42 and approached corresponding predictions from the theoretical model for CR
the distribution for CR ¼ 1:0 at low ionic strength [see of 0.42 (solid lines). Naþ concentrations and colors are as in (c).
Figs. S10 and S11(a) [17]]. Thus, the small value for CR is
specific to DNA-DNA interactions. At elevated ion con-
centrations, the mean electrostatic potential around the By combining single-molecule experiments, theoretical
DNA was too weak to create the ion distribution as ob- considerations, and coarse-grained and all-atom simula-
served in simulation and as predicted by PB theory [see tions, we have shown that, within a cylinder approximation,
Figs. S11(b) and S11(c) [17]]. This suggests that ion dis- DNA-DNA interactions can be described only by a signifi-
tributions are not only determined by electrostatics, but cantly reduced DNA charge. Furthermore, we have pro-
that other effects such as correlations in the ion clouds, the vided a theory that accurately describes DNA supercoils
noncontinuum nature of the dielectric surrounding, and ion over a broad range of tension and ionic strength. Super-
exclusion can have a significant influence. This in turn is coiling under tension is a unique way to confine and bring
likely to cause the low DNA-DNA interaction forces. two DNA molecules into close proximity in the absence of
Additionally, deviations from the homogeneously charged interfering surfaces [8]. In contrast to previous topological
rod model, such as strongly localized charges at the phos- investigations of long DNA [6,12,13], our force-based
phates and counterions entering the DNA grooves, may experiments allow a more reliable quantification of
reduce the interaction forces. DNA-DNA interactions since much smaller distances are

158101-3
week ending
PRL 105, 158101 (2010) PHYSICAL REVIEW LETTERS 8 OCTOBER 2010

a *aksiment@illinois.edu

gero.wedemann@fh-stralsund.de

ralf.seidel@biotec.tu-dresden.de
[1] D. Marenduzzo, E. Orlandini, A. Stasiak, D. W. Sumners,
L. Tubiana, and C. Micheletti, Proc. Natl. Acad. Sci.
U.S.A. 106, 22 269 (2009).
[2] A. A. Kornyshev and S. Leikin, Phys. Rev. Lett. 86, 3666
b (2001).
60 mM [3] Y. Kao-Huang, A. Revzin, A. P. Butler, P. O’Conner, D. W.
Force (pN / DNA turn)

20
170 mM Noble, and P. H. von Hippel, Proc. Natl. Acad. Sci. U.S.A.
300 mM
15
74, 4228 (1977).
[4] J. Ubbink and T. Odijk, Biophys. J. 76, 2502 (1999).
10 [5] J. F. Marko and E. D. Siggia, Phys. Rev. E 52, 2912 (1995).
[6] K. Klenin, H. Merlitz, and J. Langowski, Biophys. J. 74,
5 780 (1998).
[7] J. A. Schellman and D. Stigter, Biopolymers 16, 1415
0 (1977).
2.5 3.0 3.5 4.0 4.5
DNA-DNA separation (nm) [8] S. van Dorp, U. F. Keyser, N. H. Dekker, C. Dekker, and
S. G. Lemay, Nature Phys. 5, 347 (2009).
FIG. 4 (color). All-atom molecular dynamics simulations of [9] B. Luan and A. Aksimentiev, Phys. Rev. E 78, 021912
the effective force between double-stranded DNA. (a) All-atom (2008).
model. The DNA atoms are depicted as red spheres, counter- and [10] D. C. Rau and V. A. Parsegian, Biophys. J. 61, 246
coions as blue and purple spheres, respectively, and water as a (1992).
semitransparent molecular surface. The distance between the [11] X. Qiu, L. W. Kwok, H. Y. Park, J. S. Lamb, K. Andresen,
DNA molecules was restrained by a harmonic potential, sche- and L. Pollack, Phys. Rev. Lett. 96, 138101 (2006).
matically depicted as a spring. (b) Mean force between the DNA [12] A. Vologodskii and N. Cozzarelli, Biopolymers 35, 289
molecules from molecular dynamics simulations (circles, solid (1995).
lines) alongside the theoretical predictions for CR ¼ 0:42 [13] M. Hammermann, C. Steinmaier, H. Merlitz, U. Kapp, W.
(dashed lines) and 1.0 (dotted lines) at 60 (red) and 170 mM Waldeck, G. Chirico, and J. Langowski, Biophys. J. 73,
(black) bulk ion concentrations. Data for 300 mM (blue) are 2674 (1997).
reproduced from previous work [26]. [14] F. Mosconi, J. F. Allemand, D. Bensimon, and V.
Croquette, Phys. Rev. Lett. 102, 078301 (2009).
[15] D. Klaue and R. Seidel, Phys. Rev. Lett. 102, 028302
achieved (see Figs. S4 and S7 and discussion [17]). The (2009).
surprisingly small DNA-DNA interactions result from a [16] H. Brutzer, N. Luzzietti, D. Klaue, and R. Seidel, Biophys.
complex interplay of a highly charged and structured mole- J. 98, 1267 (2010).
cule with solvent molecules and ions. The simple effective [17] See supplementary material at http://link.aps.org/
supplemental/10.1103/PhysRevLett.105.158101 for fig-
interaction potential will be an important contribution for
ures, discussion, and methods.
quantitative models of complex biomolecular systems [18] T. R. Strick, M. N. Dessinges, G. Charvin, N. H. Dekker,
which cannot be treated with atomic detail such as DNA J. F. Allemand, D. Bensimon, and V. Croquette, Rep. Prog.
packaging in chromatin and viruses but possibly also for Phys. 66, 1 (2003).
protein-DNA interactions. The findings reveal that particu- [19] S. Forth, C. Deufel, M. Y. Sheinin, B. Daniels, J. P. Sethna,
lar caution is necessary when applying effective charge and M. D. Wang, Phys. Rev. Lett. 100, 148301 (2008).
parameters obtained from experiments that probe a differ- [20] J. F. Marko, Phys. Rev. E 76, 021926 (2007).
ent physics, such as electrophoresis [22], resulting from a [21] N. Clauvelin, B. Audoly, and S. Neukirch, Biophys. J. 96,
complex interplay between hydrodynamics and electrostat- 3716 (2009).
ics [8,9,27]. [22] D. Stigter, Biopolymers 16, 1435 (1977).
We gratefully acknowledge K. Kroy, U. Keyser, and M. [23] S. L. Brenner and V. A. Parsegian, Biophys. J. 14, 327
Emanuel for helpful discussions. This work was supported (1974).
[24] H. Zhang and J. F. Marko, Phys. Rev. E 77, 031916
by Grants No. SE 1646/1-1 and No. SE 1646/2-1 from the
(2008).
DFG to R. Seidel, the National Institutes of Health (R01- [25] A. V. Vologodskii and J. F. Marko, Biophys. J. 73, 123
HG003713 and PHS 5 P41-RR05969), the National Science (1997).
Foundation (PHY0822613), the Petroleum Research Fund [26] B. Luan and A. Aksimentiev, J. Am. Chem. Soc. 130,
(48352-G6), and the TeraGrid (MCA05S028) to A. A. as 15 754 (2008).
well as by Project No. mvb00007 of the North German [27] B. Luan and A. Aksimentiev, Soft Matter 6, 243
Supercomputing Alliance (HLRN). (2010).

158101-4

You might also like