Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

Backbone charge transport in double-stranded DNA

Abstract

Understanding charge transport in DNA molecules is a long-standing problem of fundamental importance across disciplines1,2. It is also of great technological interest due to DNA’s ability to form versatile and complex programmable structures. Charge transport in DNA-based junctions has been reported using a wide variety of set-ups2,3,4, but experiments so far have yielded seemingly contradictory results that range from insulating5,6,7,8 or semiconducting9,10 to metallic-like behaviour11. As a result, the intrinsic charge transport mechanism in molecular junction set-ups is not well understood, which is mainly due to the lack of techniques to form reproducible and stable contacts with individual long DNA molecules. Here we report charge-transport measurements through single 30-nm-long double-stranded DNA (dsDNA) molecules with an experimental set-up that enables us to address individual molecules repeatedly and to measure the current–voltage characteristics from 5 K up to room temperature. Strikingly, we observed very high currents of tens of nanoamperes, which flowed through both homogeneous and non-homogeneous base-pair sequences. The currents are fairly temperature independent in the range 5–60 K and show a power-law decrease with temperature above 60 K, which is reminiscent of charge transport in organic crystals. Moreover, we show that the presence of even a single discontinuity (‘nick’) in both strands that compose the dsDNA leads to complete suppression of the current, which suggests that the backbones mediate the long-distance conduction in dsDNA, contrary to the common wisdom in DNA electronics2,3,4.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Experimental set-up and SEM images of the dsDNA junctions.
Fig. 2: Sequence dependence of the IV characteristics.
Fig. 3: The role of the backbone in the IV characteristics.
Fig. 4: Ab initio electronic structure calculations.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding authors upon reasonable request. Source data are provided with this paper.

References

  1. Cuevas, J. C. & Scheer, E. Molecular Electronics: An Introduction to Theory and Experiment (World Scientific, 2017).

  2. Schuster, G. B. (ed.) Long-Range Charge Transfer in DNA I and II (Topics in Current Chemistry Vol 236–237, Springer, 2004).

  3. Porath, D., Cuniberti, G. & Di Felice, R. in Long-Range Charge Transfer in DNA I and II (ed. Schuster, G. B.) 183–227 (Topics in Current Chemistry Vol 236–237, Springer, 2004).

  4. Endres, R. G., Cox, D. L. & Singh, R. R. P. Colloquium: the quest for high-conductance DNA. Rev. Mod. Phys. 76, 195–214 (2004).

    CAS  Google Scholar 

  5. Braun, E., Eichen, Y., Sivan, U. & Ben-Yoseph, G. DNA-templated assembly and electrode attachment of a conducting silver wire. Nature 391, 775–778 (1998).

    CAS  Google Scholar 

  6. De Pablo, P. J. et al. Absence of dc-conductivity in λ-DNA. Phys. Rev. Lett. 85, 4992–4995 (2000).

    Google Scholar 

  7. Storm, A. J., van Noort, J., de Vries, S. & Dekker, C. Insulating behavior for DNA molecules between nanoelectrodes at the 100 nm length scale. Appl. Phys. Lett. 79, 3881–3883 (2001).

    CAS  Google Scholar 

  8. Livshits, G. I. et al. Long-range charge transport in single G-quadruplex DNA molecules. Nat. Nanotechnol. 9, 1040–1046 (2014).

    CAS  Google Scholar 

  9. Cai, L., Tabata, H. & Kawai, T. Self-assembled DNA networks and their electrical conductivity. Appl. Phys. Lett. 77, 3105–3106 (2000).

    CAS  Google Scholar 

  10. Porath, D., Bezryadin, A., de Vries, S. & Dekker, C. Direct measurement of electrical transport through DNA molecules. Nature 403, 635–638 (2000).

    CAS  Google Scholar 

  11. Fink, H. W. & Schonenberger, C. Electrical conduction through DNA molecules. Nature 398, 407–410 (1999).

    CAS  Google Scholar 

  12. Muren, N. B., Olmon, E. D. & Barton, J. K. Solution, surface, and single molecule platforms for the study of DNA-mediated charge transport. Phys. Chem. Chem. Phys. 14, 13754–13771 (2012).

    CAS  Google Scholar 

  13. Zhuravel, R. et al. Advances in synthesis and measurement of charge transport in DNA based derivatives. Adv. Mat. 30, 1706984 (2018).

    Google Scholar 

  14. Heim, T., Deresmes, D. & Vuillaume, D. Conductivity of DNA probed by conducting-atomic force microscopy: effects of contact electrode, DNA structure, and surface interactions. J. Appl. Phys. 96, 2927–2936 (2004).

    CAS  Google Scholar 

  15. Cohen, H., Nogues, C., Naaman, R. & Porath, D. Direct measurement of electrical transport through single DNA molecules of complex sequence. Proc. Natl Acad. Sci. USA 102, 11589–11593 (2005).

    CAS  Google Scholar 

  16. Xiang, L. et al. Intermediate tunnelling-hopping regime in DNA charge transport. Nat. Chem. 7, 221–225 (2015).

    CAS  Google Scholar 

  17. Artés, J. M., Li, Y., Qi, J., Anantram, M. P. & Hihath, J. Conformational gating of DNA conductance. Nat. Commun. 6, 8870 (2015).

    Google Scholar 

  18. Guo, X. F., Gorodetsky, A. A., Hone, J., Barton, J. K. & Nuckolls, C. Conductivity of a single DNA duplex bridging a carbon nanotube gap. Nat. Nanotechnol. 3, 163–167 (2008).

    CAS  Google Scholar 

  19. Dachlika, H. et al. Formation of dimers composed of a single short dsDNA connecting two gold nanoparticles. J. Self-Assem. Mol. Electron. 1, 85–99 (2013).

    CAS  Google Scholar 

  20. Pethig, R. Dielectrophoresis: status of the theory, technology, and applications. Biomicrofluidics 4, 022811 (2010).

    Google Scholar 

  21. Dadosh, T. et al. Measurement of the conductance of single conjugated molecules. Nature 436, 677–680 (2005).

    CAS  Google Scholar 

  22. Eidelshtein, G. et al. Preparation, characterization and manipulation of conjugates between gold nanoparticles and DNA. Nanomaterials 6, 167 (2016).

    Google Scholar 

  23. Bixon, M. et al. Long-range charge hopping in DNA. Proc. Natl Acad. Sci. USA 96, 11713–11716 (1999).

    CAS  Google Scholar 

  24. Beall, E. et al. Molecular conductance of nicked nucleic acid duplexes. J. Phys. Chem. C 122, 7533–7540 (2018).

    CAS  Google Scholar 

  25. Köhler, A. & Bässler, H. Electronic Processes in Organic Semiconductors: An Introduction (Wiley, 2015).

  26. Coropceanu, V. et al. Charge transport in organic semiconductors. Chem. Rev. 107, 926–952 (2007).

    CAS  Google Scholar 

  27. Motamarri, P. et al. DFT-FE: a massively parallel adaptive finite-element code for large-scale density functional theory calculations. Comput. Phys. Commun. 246, 106853 (2020).

    CAS  Google Scholar 

  28. Elstner, M. et al. Self-consistent-charge density-functional tight-binding method for simulations of complex materials properties. Phys. Rev. B 58, 7260–7268 (1998).

    CAS  Google Scholar 

  29. Kubar, T., Elstner, M., Popescu, B. & Kleinekathöfer, U. Polaron effects on charge transport through molecular wires: a multiscale approach. J. Chem. Theory Comput. 13, 286–296 (2017).

    CAS  Google Scholar 

  30. Oberhofer, Η, Reuter, Κ & Blumberger, J. Charge transport in molecular materials: an assessment of computational methods. Chem. Rev. 117, 10319–10357 (2017).

    CAS  Google Scholar 

  31. Beratan, D. N. et al. Charge transfer in dynamical biosystems, or the treachery of (static) images. Acc. Chem. Res. 48, 474–481 (2014).

    Google Scholar 

  32. Ortmann, K., Behstedt, F. & Hannewald, K. Charge transport in organic crystals: theory and modelling. Phys. Status Solidi B 248, 511–525 (2011).

    CAS  Google Scholar 

  33. Troisi, A. & Orlandi, G. Charge-transport regime of crystalline organic semiconductors: diffusion limited by thermal off-diagonal electronic disorder. Phys. Rev. Lett. 96, 086601 (2006).

    Google Scholar 

  34. Fratini, S., Mayou, D. & Ciuchi, S. The transient localization scenario for charge transport in crystalline organic materials. Adv. Funct. Mater. 26, 2292–2315 (2016).

    CAS  Google Scholar 

  35. Jana, N. R., Gearheart, L. & Murphy, C. J. Seeding growth for size control of 5−40 nm diameter gold nanoparticles. Langmuir 17, 6782–6786 (2001).

    CAS  Google Scholar 

  36. Hanwell, M. D. et al. Avogadro: An advanced semantic chemical editor, visualization and analysis platform. J. Cheminform. 4, 17 (2012).

    CAS  Google Scholar 

  37. Das, S. et al. DFT-FE: Real-Space DFT calculations Using Finite Elements https://github.com/dftfeDevelopers/dftfe (2018).

  38. Motamarri, P., Nowak, M. R., Leiter, K., Knap, J. & Gavini, V. Higher-order adaptive finite-element methods for Kohn–Sham density functional theory. J. Comput. Phys. 253, 308–343 (2013).

    CAS  Google Scholar 

  39. Das, S. et al. Fast, scalable and accurate finite-element based ab initio calculations using mixed precision computing: 46 PFLOPS simulation of a metallic dislocation system. Proc. SC19, The International Conference for High Performance Computing, Networking, Storage, and Analysis 2 (Association for Computing Machinery, 2019).

  40. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996). Erratum 78, 1396 (1997).

  41. Hamann, D. R. Optimized norm-conserving Vanderbilt pseudopotentials. Phys. Rev. B 88, 085117 (2013). Erratum 95, 239906 (2017).

  42. Brooks, B. R. et al. CHARMM: the biomolecular simulation program. J. Comput. Chem. 30, 1545–1614 (2009).

    CAS  Google Scholar 

  43. Becker, O. M., MacKerell, A. D., Roux, B. & Watanabe, M. Computational Biochemistry and Biophysics (Marcel Dekker, 2001).

  44. Ryckaert, J. P., Ciccotti, G. & Berendsen, H. J. C. Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comput. Phys. 23, 327–341 (1977).

    CAS  Google Scholar 

  45. Darden, T., York, D. & Pedersen, L. Particle mesh Ewald: an Nlog(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    CAS  Google Scholar 

  46. Hoover, W. G. Canonical dynamics: equilibrium phase-space distributions. Phys. Rev. A 31, 1695–1697 (1985).

    CAS  Google Scholar 

  47. Mizoguchi, K., Tanaka, S., Ogawa, T., Shiobara, N. & Sakamoto, H. Magnetic study of the electronic states of B-DNA and M-DNA doped with metal ions. Phys. Rev. B 72, 033106 (2005).

    Google Scholar 

  48. Kino, H. et al. A possible origin of carrier doping into DNA. J. Phys. Soc. Jpn 73, 2089–2092 (2004).

    CAS  Google Scholar 

  49. Gervasio, F. L., Laio, A., Parrinello, M. & Boero, M. Charge localization in DNA fibers. Phys. Rev. Lett. 94, 158103 (2005).

    Google Scholar 

  50. Aradi, B. & Hourahine, B. & Frauenheim., Th. DFTB+, a sparse matrix-based implementation of the DFTB method. J. Phys. Chem. A 111, 5678–5684 (2007).

    CAS  Google Scholar 

  51. Kubas, A. et al. Electronic couplings for molecular charge transfer: benchmarking CDFT, FODFT, FODFTB against high-level ab initio calculations. J. Chem. Phys. 140, 104105 (2014).

    Google Scholar 

Download references

Acknowledgements

This research was supported by the Israel Science Foundation (ISF grant nos. 1589/14 and 2556/17) and by the Minerva Centre for Bio-Hybrid Complex Systems. D.P. thanks the Etta and Paul Schankerman Chair of Molecular Biomedicine. G.P. and S.S.S. acknowledge financial support from the University of Cyprus PhD grants and J.C.C. from the Spanish MINECO (contract no. FIS2017-84057-P). V.G. acknowledges the support of Toyota Research Institute under the auspices of which some aspects of the DFT-FE package relevant to this work were developed. L.A.Z. thanks financial support from the University of Seville through the VI PPIT-US program. This paper and work are dedicated to the memory of Professor Joseph Sperling, who passed away during the performance of this research.

Author information

Authors and Affiliations

Authors

Contributions

R.Z., D.R. and D.P. conceived the work. R.Z. prepared the dimers based on inspiration and guidance by J.S. R.Z. and H.H. fabricated the devices. R.Z. performed the transport experiments and analysed the data. A.B.K. and L.K. prepared the poly(dG)–poly(dC) dimers. G.P., S.P., P.M. and L.A.Z. performed the calculations under the supervision of S.S.S., V.G. and J.C.C. R.Z., A.B.K., J.C.C., S.S.S., D.R. and D.P. wrote the manuscript, with comments and input from all the authors.

Corresponding author

Correspondence to Danny Porath.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Dimer synthesis.

a, Scheme of dimer synthesis. Incubation of GNPs and ssDNA end-modified with thiol followed by purification of the conjugates using electrophoresis yields the mono-conjugate composed of one DNA oligonucleotide attached to the GNP. Mixing the mono-conjugates comprising complementary DNA strands results in formation of dimers. b, TEM scan of the dimer solution. Typically, 40–90% of the GNPs were found in dimers.

Extended Data Fig. 2 Temperature dependence of the current for different dsDNA molecules.

a, Arrhenius plot for a random sequence dsDNA molecule that was heated from 5 K to 125 K and cooled down back to 5 K. The transition is reversible. b, Arrhenius plot of different molecules (random sequence, with and without nicks and poly(dG)-poly(dC)). All the dsDNA molecule types (see legend) have a transition temperature at 40-80 K with nearly constant current below the transition and rapid decrease above it.

Source data

Extended Data Fig. 3 2D histograms of I-V curves for each type of DNA.

Two-dimensional (2D) histograms of I-V measurements at temperatures below the transition temperature for each type of molecule. For clarity, only data points that appear in more than one measurement are presented here. The histograms show that I-V curves from different molecules of the same type are generally similar but not identical due to the variability in the contacts and positioning. Nevertheless, the different types of dsDNA molecules, share similar trends. a, 100 base-pairs random sequence dsDNA: 217 curves from 13 molecules. b, 100 base pairs Poly(dG)-poly(dC): 43 curves from 6 molecules. c, The single-nicked DNA having a sequence identical to that in (a): 22 curves from 3 molecules.

Extended Data Fig. 4 Ab initio electronic structure calculations of B-type dsDNA molecules.

a, Highest occupied molecular orbital (HOMO), which resides in the base pairs, and the highest occupied orbital that belongs to backbone atoms for a B-type 100 base-pairs poly(dG)-poly(dC) dsDNA molecule. The energy of these two orbitals, with respect to the vacuum energy, is indicated in this panel. Notice in both cases the localized nature of these orbitals. b, The corresponding local density of states (LDOS) as a function of energy for the poly(dG)-poly(dC) dsDNA molecule showing the individual contributions of base-pair-centred states (black line) and backbone-centred states (red line). c, The computed localization length, normalized by the length of the molecule, as a function of energy for the occupied states of the poly(dG)-poly(dC) dsDNA molecule. The vertical dashed lines in panels b and c indicate the position of the HOMO. d-f, The same as in panels a-c for a B-type 100 base-pairs dsDNA molecule with the random sequence investigated experimentally. All the results were obtained using density functional theory and the DFT-FE package27.

Source data

Extended Data Fig. 5 Ab initio electronic structure calculations of nicked molecules.

a, A-type 100 base-pairs dsDNA molecule with a random sequence and featuring a nick (removal of a phosphate group in the position indicated in the frame). b, The same molecule as in panel a, but featuring two nicks (one in each strand, as indicated by the two frames). c, The local density of states (LDOS) as a function of energy for the one-nick molecule of panel a showing the individual contributions of base-pair-centred states (black line) and backbone-centred states (red line). d, The computed localization length, normalized by the total length of the molecule, as a function of energy for the occupied states of the one-nick molecule of panel a. The vertical dashed lines in panels c and d indicate the position of the HOMO. d-f, The same as in panels c-d for the two-nick molecule of panel b. All the results were obtained using density functional theory and the DFT-FE package27.

Source data

Extended Data Fig. 6 Molecular dynamics simulations of 8 base-pairs dsDNA molecules.

a, A hydrogen bond occurs when a hydrogen atom (H) and the acceptor (A) are closer than 2.4Å and the angle between D – H – A is greater than 140°. b, Schematic drawing of Watson-Crick pairs guanine:cytosine. c, The distance between neighboring nucleobases is specified by the vector connecting their centers of mass. d, The vectors \(\vec a_i,\vec b_i\) define a plane for the nucleobase and the normal to the plane is described by \(\vec c_i = \vec a_i \times \vec b_i\). The relative orientation between adjacent bases is described by \(cos\theta = \frac{{\vec c_i \cdot \vec c_{i + 1}}}{{\left| {\vec c_i} \right|\left| {\vec c_{i + 1}} \right|}}\). e,f The figures show the relative position and orientation between neighboring nucleobases in the same strand only for the bases around the nicks. e, Intact structures, poly(dG)-poly(dC) and random sequence. f, One-nicked structures, poly(dG)-poly(dC) and random sequences.

Extended Data Fig. 7 Molecular dynamics simulations of intact and nicked random sequence dsDNA molecules.

a, A-type 8-base-pairs dsDNA molecule with a random sequence. The histograms show the probability distributions of inter-base distances and angles (Extended Data Fig. 6 and Supplementary Table 8) for the base pair along a single strand that is adjacent to the phosphate group to be removed in the nicked structure. b, Probability distributions as in a for the same molecule with a nick (removal of a phosphate group) inserted between the indicated base pairs. For notation see Supplementary Table 3. The MD simulations were carried out at 300 K using the CHARMM c37b2 software package42.

Source data

Extended Data Fig. 8 Molecular dynamics simulations and DFTB electronic structure computations of intact and nicked random sequence dsDNA molecules.

A-type 8-base-pairs dsDNA molecule with a random sequence. a Ratio of 〈LL(E)〉nicked/〈LL(E)〉without nicks where \(\cdots\) denotes MD averaging (over 10 ns trajectories) and LL(E) denotes localization length for orbitals at energy E. b Ratios σ/〈LL(E)〉 for the three structures, where \(\sigma = \sqrt {(LL\left( E \right) - \langle LL(E)\rangle)^2}\), for the intact and nicked structures. The calculations were carried out using the Density Functional based Tight Binding (DFTB)28 method as implemented in the DFTB+50 program package (version 17.1).

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1–8.

Supplementary Data 1

Raw data – random sequence.

Supplementary Data 2

Raw data – random sequence.

Supplementary Data 3

Raw data – random sequence.

Supplementary Data 4

Raw data – random sequence.

Supplementary Data 5

Raw data – random sequence.

Supplementary Data 6

Raw data – random sequence.

Supplementary Data 7

Raw data – random sequence.

Supplementary Data 8

Raw data – random sequence.

Supplementary Data 9

Raw data – random sequence.

Supplementary Data 10

Raw data – random sequence.

Supplementary Data 11

Raw data – random sequence.

Supplementary Data 12

Raw data – random sequence.

Supplementary Data 13

Raw data – random sequence.

Supplementary Data 14

Raw data – random sequence.

Supplementary Data 15

Raw data – one nick.

Supplementary Data 16

Raw data – one nick.

Supplementary Data 17

Raw data – one nick.

Supplementary Data 18

Raw data – poly(G)–poly(C).

Supplementary Data 19

Raw data – poly(G)–poly(C).

Supplementary Data 20

Raw data – poly(G)-poly(C).

Supplementary Data 21

Raw data – poly(G)-poly(C).

Supplementary Data 22

Raw data – poly(G)-poly(C).

Supplementary Data 23

Raw data – poly(G)-poly(C).

Source data

Source Data Fig. 2

Numerical data used to generate graphs in Fig. 2a–c

Source Data Fig. 3

Numerical data used to generate graphs in 3b

Source Data Fig. 4

Numerical data used to generate graphs in 4b,c,e,f

Source Data Extended Data Fig. 2

Numerical data used to generate graphs in Extended Data Fig. 2a,b

Source Data Extended Data Fig. 4

Numerical data used to generate graphs in Extended Data Fig. 4b,c,e,f

Source Data Extended Data Fig. 5

Numerical data used to generate graphs in Extended Data Fig. 5–f

Source Data Extended Data Fig. 7

Numerical data used to generate graphs in Extended Data Fig. 7a,b

Source Data Extended Data Fig. 8

Numerical data used to generate graphs in Extended Data Fig. 8a,b

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhuravel, R., Huang, H., Polycarpou, G. et al. Backbone charge transport in double-stranded DNA. Nat. Nanotechnol. 15, 836–840 (2020). https://doi.org/10.1038/s41565-020-0741-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41565-020-0741-2

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing