(810b) Confinement of a DNA Dodecamer Onto Pristine Carbon Nanotubes: Stability of the Canonical B Form | AIChE

(810b) Confinement of a DNA Dodecamer Onto Pristine Carbon Nanotubes: Stability of the Canonical B Form

Authors 

Cruz, F. J. A. L. - Presenter, Universidade Nova de Lisboa
Mota, J. P. B., Universidade Nova de Lisboa
de Pablo, J. J., University of Chicago



Confinement of a DNA Dodecamer onto Pristine Carbon
Nanotubes and the Stability of the Canonical B Form

Fernando
J.A.L. Cruz1,2, Jose P. B. Mota1
and Juan J. de Pablo2,3

1 Requimte/CQFB, Dept. Chemistry, Universidade Nova de Lisboa, Caparica, Portugal. 2 Dept.
Chem. Bio. Eng., University of Wisconsin-Madison, Madison, WI, USA. 3 Institute of
Molecular Engineering, University of Chicago, Chicago, Illinois, USA.

fj.cruz@fct.unl.pt

I. Introduction

     Deoxyribonucleic acid
(DNA) and single-walled carbon nanotubes (SWCNTs) are prototypical
one-dimensional structures; the former plays a central role in chemical biology
and the latter holds promise in nanotechnology applications.1-3 From the point of view of biological
purposes and DNA manipulation, carbon nanotubes have been proposed as templates
for DNA encapsulation, intracellular penetration via endocytosis and delivery
of biological payloads and ultrafast nucleotide sequencing. While structure in
its natural form and environment is well established (e.g. B-DNA in aqueous solution), their interactions have been the
subject of intense investigation, nonetheless, the corresponding
molecular-level phenomena remain rather unexplored. Is confinement spontaneous
from a thermodynamical point of view (free-energy)? How important are the
conformational properties of the double strand upon confinement (entropy)? What
is the effect of the media electrical charge density on the negatively charged
DNA backbone (ionic strength)? How does the confinement process depend on intrinsic
properties of the nanotube such as diameter (D)? Previous theoretical and experimental work has focused almost
exclusively on exoadsorption of DNA at the SWCNTs external walls, overlooking
the possibility of endohedral confinement; the energetics of the latter is one
of the main purposes of the present work. Confinement driven adsorption is a
phenomenon that is utterly unmapped, and most of the previous work has been
focused in systems at temperatures away from the physiological value and/or
ionic strength, thus precluding extrapolation of results to in vivo conditions. We'll address this
issue using molecular dynamics and enhanced sampling techniques (metadynamics,
umbrella sampling) to probe the encapsulation mechanism of an atomistically
detailed DNA dodecamer, onto SWCNTs of different diameters, employing precise
physiological conditions, 310 K and [NaCl] = 134 mM.

II.
Results and Discussion

     Confinement
of DNA onto a (51,0) topology, D =
4nm, is fast and becomes complete before 16 ns of observation time (Fig.1A);
initially (t = 0 – 2 ns), the
double strand is in the bulk, and as it diffuses towards the SWCNT undergoes
structural rearrangements leading to minor increases in pitch length, P, and end-to-end distance, L. After 2 ns the dodecamer is already
at the nanopore entrance, experiencing strong van der Waals attractions towards
the solid, leading to complete encapsulation at 15.42 ns, after which it
relaxes and reassumes its unconfined structure (P = 3.4 nm, L = 4.1 nm)
consistent with the geometrical characteristics of B-DNA in bulk.4 It is interesting to observe that
confinement is unalterable, and the biomolecule never returns back to the bulk
solution during the observation time window, always maintaining direct contact
with the solid at a distance of closest approach of ca. 0.26 nm; nonetheless, the confined molecule retains its
translational mobility diffusing along the nanopore main axis (Fig. 1B). In
contrast, the narrowness of a (40,0) topology, D = 3 nm, compared to the DNA skeletal diameter, inhibits
encapsulation even for an observation time of 0.1 μs.
We determined the characteristic lengths, radius of gyration (Rg)
and its projection along the nanotube main axis (
), as well as the root-mean squared deviation (RMSD) compared to the B-DNA dodecamer
used as starting configuration; results are plotted in Fig. 1C. Because the RMSD compares the structure at any time t with the original DNA structure (t = 0), the blue curve recorded in Fig.
1C suggests a minor relaxation of the double-strand from the crystal structure
to accommodate liquid state flexibility, whilst maintaining the relative average
distance between each atom in the double-strand. After 20 ns, results have
smoothly converged to average values of RMSD
= 0.36 ± 1.9 ×10–3 nm and
 = 1.02 ± 1.5
×10–3 nm, with the latter indicating an alignment of the
biomolecule with the nanopore main axis. The canonical Watson-Crick H-bonds
network is roughly maintained throughout confinement, exhibiting probability
distributions corresponding to more than 75 %; the
 = 0.8 nm
depression observed at 56.6 ns is transient and matches a total number of
canonical H bonds of 28.

     The
thermodynamical stability of encapsulation is probed by the free-energy (F) differences associated with the
process, using well-tempered metadynamics.5
An inspection of the resulting three-dimensional surface (Fig.2) reveals the
existence of five distinct free-energy minima, sharing in common the fact that
all are located at discrete positions along the nanopore internal volume, ξ1 < 1.8 nm; the
absolute minimum at ξ1
= 0.117 nm evidences the SWCNT center as the energetically most stable region
subsequent to encapsulation. In order to escape from those deep free-energy
valleys, F ~ – 40 kJ/mol–1,
the DNA molecule has to overcome large energetic barriers, rendering the exit
process towards bulk solution a thermodynamically expensive one. The number
density maps depicted in Fig.3, obtained from atomically detailed mass
histograms, corroborate the finding that free-energy is minimum at the nanopore
center; Fig.3 also reveals a thin and cylindrically-shaped exclusion volume
close to the hydrophobic solid walls, arising from the repulsion between the
latter and the heavily charged phosphate groups. It is remarkable to observe
that the exact position of the free-energy minima oscillate little around ξ2 = 4.1 nm, coherent with L = 4.1 nm characteristic of a
B–DNA structure (Fig.1).

     As
far as we are aware these observations are the first of their kind, and they
come to pave the way for the design of smart nanotube based devices for in vivo DNA encapsulation.

FIGURE 1 - Encapsulation of DNA onto a (51,0) SWCNT. A) Kinetics. Encapsulation is complete after 15.42 ns and the double-strand never returns back to the bulk solution during the observation time window; DNA maintains its translational mobility within the nanopore. Notice that a 2 nm c.o.m. distance between the biomolecule and the solid corresponds to the threshold bellow which complete encapsulation occurs: black) distance between centres of mass of DNA and (51,0) SWCNT, dark red) distance between centres of mass of DNA and (40,0) SWCNT, green) DNA end-to-end length, blue) DNA pitch length, grey) minimum distance between any DNA atom and the SWCNT. B) State diagram. Distances between terminal nucleobase pairs and the SWCNT entrance, projected along the nanotube main axis: black) terminus 1 and red) terminus 2. The horizontal dashed lines correspond to the SWCNT boundaries and the inset magnification depicts the first instant just after complete encapsulation, representing the terminus 1 atoms in black and the terminus 2 atoms in red. The encapsulation mechanism can be described by a three-step kinetics: I) fast diffusion of bulk DNA towards the nanopore (0 – 2 ns), II) strong van der Waals attraction towards the solid, leading to confinement of terminus 1 at 2.33 ns, and finally III) penetration of terminus 2 onto the nanopore volume resulting in complete encapsulation of DNA. Note the confined molecule maintains its translational mobility along the nanopore axis. C) Characteristic lengths. black) radius of gyration, Rg, grey) radius of gyration z-component, , and blue) root mean-squared deviation, RMSD. After confinement, molecular configuration deviates minimally from the canonical B-DNA form, RMSD = 0.36 ± 1.9 ×10–3 nm, and maintains a quasi-linearity about the nanopore main axis.

FIGURE 2 - Free energy landscape of DNA@(51,0)SWCNT hybrid. ξ1 is the distance between centres of mass of the DNA and SWCNT, projected along the nanopore main axis, and ξ2 is the absolute distance between (GC) termini on opposite sides of the double-strand, equivalent to the DNA end-to-end length. The several free-energy minima along ξ1 demonstrate that the molecule is relatively mobile to translocate along the nanotube, however, the absolute minimum at ξ1 = 0.117 nm indicates that the nanopore center is the energetically favoured region. Interestingly, all the ξ1 minima are located along a quasi-linear path defined by ξ2 Å 4.1 nm highlighting the enhanced thermodynamical stability corresponding to the canonical B form. Snapshots were taken at (ξ1, ξ2) nm = A) (0.117, 4.112), B) (0.621, 4.164), C) (1.307, 4.164) and D) (1.796, 4.115).

FIGURE 3 – Density maps of DNA@(51,0)SWCNT. The existence of a cylindrical exclusion volume close to the walls,  r Å 0, is the direct consequence of direct repulsion between the heavily charged phosphate groups and the hydrophobic solid. The dashed lines indicate the boundaries of the nanotube.

References

     (1)    H. Kumar et al, Soft Matter 2011, 7, 5898.

     (2)    B.M. Venkatesan, R. Bashir, Nature Nano. 2011, 6.

     (3)    A. D. Franklin et al, Nano Lett. 2012, 12, 758-762.

     (4)    J.M. Vargason et al, Proc. Nat. Acad. Sci. 2001,
98, 7265.

     (5)    A. Barducci et al, Phys. Rev. Let. 2008, 100, 020603.