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:

RETRACTED ARTICLE: Living annulative π-extension polymerization for graphene nanoribbon synthesis

A Retraction to this article was published on 25 November 2020

An Addendum to this article was published on 18 September 2020

This article has been updated

Abstract

The properties of graphene nanoribbons (GNRs)1,2,3,4,5—such as conductivity or semiconductivity, charge mobility and on/off ratio—depend greatly on their width, length and edge structure. Existing bottom–up methods used to synthesize GNRs cannot achieve control over all three of these parameters simultaneously, and length control is particularly challenging because of the nature of step-growth polymerization6,7,8,9,10,11,12,13,14,15,16,17,18. Here we describe a living annulative π-extension (APEX)19 polymerization technique that enables rapid and modular synthesis of GNRs, as well as control over their width, edge structure and length. In the presence of palladium/silver salts, o-chloranil and an initiator (phenanthrene or diphenylacetylene), the benzonaphthosilole monomer polymerizes in an annulative manner to furnish fjord-type GNRs. The length of these GNRs can be controlled by simply changing the initiator-to-monomer ratio, achieving the synthesis of GNR block copolymers. This method represents a type of direct C–H arylation polymerization20 and ladder polymerization21, activating two C–H bonds of polycyclic aromatic hydrocarbons and constructing one fused aromatic ring per chain propagation step.

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

Access options

Buy this article

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

Fig. 1: APEX polymerization for structurally well defined GNRs.
Fig. 2: STM and AFM images of fjord-type GNR 2.
Fig. 3: Living APEX block copolymerization and highly controlled living APEX polymerization.
Fig. 4: Transformation of fjord-type GNR 2 to armchair-type GNR 8.

Similar content being viewed by others

Data availability

The datasets generated and/or analysed during the current study are available from the corresponding authors on reasonable request.

Change history

References

  1. Dutta, S. & Pati, S. K. Novel properties of graphene nanoribbons: a review. J. Mater. Chem. 20, 8207–8223 (2010).

    Article  CAS  Google Scholar 

  2. Li, X., Wang, X., Zhang, L., Lee, S. & Dai, H. Chemically derived, ultrasmooth graphene nanoribbon semiconductors. Science 319, 1229–1232 (2008).

    Article  ADS  CAS  Google Scholar 

  3. Son, Y. W., Cohen, M. L. & Louie, S. G. Half-metallic graphene nanoribbons. Nature 444, 347–349 (2006); corrigendum 446, 342 (2007).

    Article  ADS  CAS  Google Scholar 

  4. Yang, L., Park, C. H., Son, Y. W., Cohen, M. L. & Louie, S. G. Quasiparticle energies and band gaps in graphene nanoribbons. Phys. Rev. Lett. 99, 186801 (2007).

    Article  ADS  Google Scholar 

  5. Ritter, K. A. & Lyding, J. W. The influence of edge structure on the electronic properties of graphene quantum dots and nanoribbons. Nat. Mater. 8, 235–242 (2009).

    Article  ADS  CAS  Google Scholar 

  6. Ma, L., Wang, J. & Ding, F. Recent progress and challenges in graphene nanoribbon synthesis. ChemPhysChem 14, 47–54 (2013).

    Article  CAS  Google Scholar 

  7. Segawa, Y., Ito, H. & Itami, K. Structurally uniform and atomically precise carbon nanostructures. Nat. Rev. Mater. 1, 15002 (2016).

    Article  ADS  CAS  Google Scholar 

  8. Chen, L., Hernandez, Y., Feng, X. & Müllen, K. From nanographene and graphene nanoribbons to graphene sheets: chemical synthesis. Angew. Chem. Int. Ed. 51, 7640–7654 (2012).

    Article  CAS  Google Scholar 

  9. Narita, A., Feng, X. & Müllen, K. Bottom-up synthesis of chemically precise graphene nanoribbons. Chem. Rec. 15, 295–309 (2015).

    Article  CAS  Google Scholar 

  10. Yang, X. et al. Two-dimensional graphene nanoribbons. J. Am. Chem. Soc. 130, 4216–4217 (2008).

    Article  CAS  Google Scholar 

  11. Cai, J. et al. Atomically precise bottom-up fabrication of graphene nanoribbons. Nature 466, 470–473 (2010).

    Article  ADS  CAS  Google Scholar 

  12. Narita, A. et al. Synthesis of structurally well-defined and liquid-phase-processable graphene nanoribbons. Nat. Chem. 6, 126–132 (2014).

    Article  CAS  Google Scholar 

  13. Ruffieux, P. et al. On-surface synthesis of graphene nanoribbons with zigzag edge topology. Nature 531, 489–492 (2016).

    Article  ADS  CAS  Google Scholar 

  14. Talirz, L. et al. On-surface synthesis and characterization of 9-atom wide armchair graphene nanoribbons. ACS Nano 11, 1380–1388 (2017).

    Article  CAS  Google Scholar 

  15. Sakaguchi, H., Song, S., Kojima, T. & Nakae, T. Homochiral polymerization-driven selective growth of graphene nanoribbons. Nat. Chem. 9, 57–63 (2017).

    Article  CAS  Google Scholar 

  16. Jordan, R. S. et al. Synthesis of graphene nanoribbons via the topochemical polymerization and subsequent aromatization of a diacetylene precursor. Chem 1, 78–90 (2016).

    Article  CAS  Google Scholar 

  17. Jordan, R. S. et al. Synthesis of N = 8 armchair graphene nanoribbons from four distinct polydiacetylenes. J. Am. Chem. Soc. 139, 15878–15890 (2017).

    Article  CAS  Google Scholar 

  18. Yang, W., Lucotti, A., Tommasini, M. & Chalifoux, W. A. Bottom-up synthesis of soluble and narrow graphene nanoribbons using alkyne benzannulations. J. Am. Chem. Soc. 138, 9137–9144 (2016).

    Article  CAS  Google Scholar 

  19. Ito, H., Ozaki, K. & Itami, K. Annulative π-extension (APEX): an enabling reaction for rapid access to fused aromatics, heteroaromatics, and nanographenes. Angew. Chem. Int. Ed. 56, 11144–11164 (2017).

    Article  CAS  Google Scholar 

  20. Pouliot, J.-R., Grenier, F., Blaskovits, J. T., Beaupré, S. & Leclerc, M. Direct (hetero)arylation polymerization: simplicity for conjugated polymer synthesis. Chem. Rev. 116, 14225–14274 (2016).

    Article  CAS  Google Scholar 

  21. Teo, Y. C., Lai, H. W. H. & Xia, Y. Synthesis of ladder polymers: developments, challenges, and opportunities. Chemistry 23, 14101–14112 (2017).

    Article  CAS  Google Scholar 

  22. Ozaki, K., Kawasumi, K., Shibata, M., Ito, H. & Itami, K. One-shot K-region-selective annulative π-extension for nanographene synthesis and functionalization. Nat. Commun. 6, 6251 (2015).

    Article  ADS  CAS  Google Scholar 

  23. Yano, Y., Ito, H., Segawa, Y. & Itami, K. Helically twisted tetracene: synthesis, crystal structure, and photophysical properties of hexabenzo[a,c,fg,j,l,op]tetracene. Synlett 27, 2081–2084 (2016).

    Article  CAS  Google Scholar 

  24. Yokozawa, T. & Ohta, Y. Transformation of step-growth polymerization into living chain-growth polymerization. Chem. Rev. 116, 1950–1968 (2016).

    Article  CAS  Google Scholar 

  25. Huang, Y. et al. Poly(ethylene oxide) functionalized graphene nanoribbons with excellent solution processability. J. Am. Chem. Soc. 138, 10136–10139 (2016).

    Article  CAS  Google Scholar 

  26. Webster, O. W. Living polymerization methods. Science 251, 887–893 (1991).

    Article  ADS  CAS  Google Scholar 

  27. Ozaki, K. et al. One-step annulative π-extension of alkynes with dibenzosiloles or dibenzogermoles by palladium/o-chloranil catalysis. Angew. Chem. Int. Ed. 56, 1361–1364 (2017).

    Article  CAS  Google Scholar 

  28. Prezzi, D., Varsano, D., Ruini, A., Marini, A. & Molinari, E. Optical properties of graphene nanoribbons: the role of many-body effects. Phys. Rev. B 77, 041404 (2008).

    Article  ADS  Google Scholar 

  29. Vandescuren, M., Hermet, P., Meunier, V., Henrard, L. & Lambin, P. Theoretical study of the vibrational edge modes in graphene nanoribbons. Phys. Rev. B 78, 195401 (2008).

    Article  ADS  Google Scholar 

  30. Gillen, R., Mohr, M. & Maultzschm, J. Symmetry properties of vibrational modes in graphene nanoribbons. Phys. Rev. B 81, 205426 (2010).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This work was supported by the ERATO programme from JST (JPMJER1302 to K.I.), the JSPS KAKENHI (grants JP26810057, JP16H00907, JP17K1955 and JP18H02019 to H.I.), the SUMITOMO Foundation (141495 to H.I.) and the DAIKO Foundation (H.I.). We acknowledge Taoka Chemical Co., Ltd for providing samples. We thank K. Kuwata (Nagoya University) for MALDI–TOF mass measurements; M. Kamigaito and M. Uchiyama (Nagoya University) for measurements of molecular weight by size-exclusion chromatography multi-angle light scattering (SEC-MALS); T. Hashizume (Hitachi Ltd and Tokyo Institute of Technology) for advice on the STM and AFM measurements; C. M. Crudden (Queen’s University and Nagoya University), G. J. P. Perry (Nagoya University) and A. Miyazaki (Nagoya University) for comments and proofreading. Computations were performed at the Research Center for Computational Science, Okazaki, Japan. ITbM is supported by the World Premier International Research Center (WPI) Initiative, Japan.

Reviewer information

Nature thanks Lawrence T. Scott and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

K.I. and H.I. conceived the idea and directed the project. Y.Y., K. Matsushima and H.I. conducted the experiments and theoretical calculations. K. Matsui, A.T. and N.M. conducted the STM experiments. F.W. conducted the AFM experiments. Y.M. supervised the scanning probe microscopy experiments and provided advice on spectroscopic analysis. H.I. and K.I. prepared the manuscript with feedback from the other authors.

Corresponding authors

Correspondence to Hideto Ito or Kenichiro Itami.

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 figures and tables

Extended Data Fig. 1 Mass spectroscopy analysis and confirmation of livingness of APEX polymerization.

a, MALDI–TOF mass spectrum (reflection mode) of fjord-type GNR 2 (M n = 1.3 × 104, Đ = 1.23) with trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylidene]malononitrile (DCTB) as the matrix. b, Summary of observed and calculated molecular ion peaks. c, Confirmation of livingness of APEX polymerization by one-pot SEC analyses. d, SEC charts of the reaction progress after 1 h, 3 h and 12 h.

Extended Data Fig. 2 Raman spectroscopy analyses of fjord-type GNR 2 and synthesis of model dimer.

a, Raman spectra of fjord-type GNR 2, obtained with a 532-nm excitation laser. b, Synthesis of model dimer 3b by step-by-step APEX reaction of phenanthrene. c, Synthesis of model dimer 3a by APEX reaction.

Extended Data Fig. 3 1H NMR spectroscopy analysis of fjord-type GNR 2.

Observed 1H NMR spectra of GNR 2 (in Cl2CDCDCl2 at 120 °C) and dimer 3a (in CD2Cl2 at 22 °C). All of the peaks of 3a range from 7.2 to 7.9 p.p.m. Two doublet-of-doublet peaks appear at a higher magnetic field (chemical shift δ = 7.2–7.3 p.p.m.), which are attributed to the most shielded aromatic hydrogen atoms, in the orange-coloured fjord regions. Other aromatic hydrogen atoms in the fjord regions next to the alkyl substituents (red circle) are found as a singlet peak at 7.32 p.p.m. On the other hand, the most deshielded aromatic hydrogen atoms in the bay and cove regions appear at 7.8–7.9 p.p.m. The 1H NMR spectrum of fjord-type GNR 2 shows broadened peaks at 7.0–8.3 p.p.m., which is expected to reflect the repeating [5]helicene structure.

Extended Data Fig. 4 Comparison of IR, absorption, emission and Raman spectroscopy of 2 and 5.

a, Observed IR spectra of GNRs 2 and 5 (blue and red lines) and IR spectrum of tetramer 10 (black line) predicted by DFT calculations at the B3LYP/6-31G(d) level of theory. b, UV-vis absorption spectra of GNRs 2 and 5 (blue and red lines). c, Emission spectra of GNRs 2 and 5 (blue and red lines) excited at 320 nm. d, Raman spectra of GNR 5, obtained by excitation by a 532-nm laser.

Extended Data Fig. 5 Analytical SEC, IR and Raman spectroscopy data for fjord-type GNR 7.

a, SEC charts of the crude product of reactions with various I′/M ratios. b, c, M n and M w profiles versus M/I′ and conversion of the monomer in the living APEX polymerization. d, Observed IR spectra of GNR 7 with different polymer lengths (blue, green, yellow and red lines) and IR spectrum of dimer 11 (black line) predicted by DFT calculations at the B3LYP/6-31G(d) level of theory. e, Raman spectra of GNR 7 with different polymer lengths (blue, green, yellow and red lines), obtained with excitation by a 532 nm laser.

Extended Data Fig. 6 Analytical MALDI–TOF data for fjord-type GNR 7.

a, MALDI–TOF mass spectrum (reflection mode) of fjord-type GNR 7 (M n = 1.2 × 104 Da, Đ = 1.04) with DCTB as the matrix. b, Summary of observed and calculated molecular ion peaks.

Extended Data Fig. 7 Formation of armchair-type GNR 8.

ad, MALDI–TOF mass spectra (reflection mode) of armchair-type GNR 8 (M n = 8.3 × 103 Da, Đ = 1.31) with 7,7,8,8-tetracyanoquinodimethane as the matrix and AgOCOCF3 as an additive. e, Summary of observed and calculated molecular ion peaks. The formation of GNR 8 with M n = 2.4 × 104 Da (Đ = 1.32) was confirmed by MALDI–TOF mass spectroscopy analysis, where 11 distinct mass peaks, each with gaps of m/z = 378 (DP(n) = 21–31), were observed. Around the [M+Ag]+ peaks with n = 25 and 27, mass peaks derived from two additional hydrogen atoms ([M+Ag+2H]+) were clearly observed, which indicates partial completion of the Scholl reaction. f, Efficiency of the transformation reaction to the armchair-type GNR 8 10. According to the literature10, we also estimated the efficiency of cyclodehydrogenation at each observed mass peak by using the theoretical number of lost hydrogen atoms during cyclodehydrogenation (N H(T.)) and the number of lost hydrogen atoms derived from MALDI–TOF mass measurements (N H(R.)) As a result, 96%–100% efficiencies (N H(T.)/N H(R.)) were confirmed at each m/z peak. The highest peak was found at m/z = 10,121.3213 (DP(n) = 26, [M+107Ag]+), which is higher than the PS-based M n, but the top peak DP (DPtop) and the dispersity pattern were almost identical to those of the starting fjord-type GNR 2. No peak corresponding to shorter oligomers or fragments was observed in the full-scale MALDI–TOF mass spectrum (Supplementary Fig. 19). The detection of larger GNRs with a molecular weight of over 20,000 Da appeared to be difficult owing to limitations in the measurement of the highly aggregating flat armchair GNR in the MALDI–TOF mass instrument.

Extended Data Fig. 8 Analytical SEC, UV-vis absorption, emission, IR and Raman spectroscopy data for armchair-type GNR 8.

a, SEC charts of armchair-type GNR 8 obtained by the Scholl reaction of fjord-type GNR 2 with different polymer lengths. b, UV-vis absorption spectra of 8. c, Emission spectra of 8. d, Observed IR spectra of GNR 8 with different polymer lengths (red, yellow, blue and purple lines) and IR spectrum of dimer 11 (black line) predicted by DFT calculations at the B3LYP/6-31G(d) level of theory. e, Raman spectra of GNR 8 with different polymer lengths (red, yellow, blue and purple lines), obtained by excitation by a 532-nm laser.

Extended Data Fig. 9 STM and AFM images of armchair-type GNR 8.

a, STM image of armchair-type GNR 8 (M n = 2.4 × 104 Da, Đ = 1.32) deposited on HOPG (U = 1.0 V, I = 250 pA, T = 78 K). b, Dimeric assembly model of GNR 8. c, d, Cross-sectional height profile taken perpendicularly (i) and horizontally (ii) with respect to the wire shown in a. The observed substance had about 0.4 nm thickness and 3 nm width, and the repeating peripheral stripes had 0.4–0.5 nm intervals ((i) and (ii)). The existence of this substance was considered as the assembly of two molecules of 8, as depicted in b, and the observed periodicity of 0.4–0.5 nm corresponds to the longitudinal length between the alkyl side chains ((ii) in d). e, AFM phase image of self-assembling GNR 8 (M n = 2.4 × 104 Da, Đ = 1.32) on graphene cleaved on SiO2. f, Magnified AFM phase image of e. g, Cross-sectional height profile taken along the white line in f, perpendicularly to the observed stripe.

Supplementary information

Supplementary Information

This file contains preparation methods for all compounds, methods for spectroscopic analysis and microscopic observations, DFT calculations, Supplementary Figures 1–42, Supplementary Tables 1 and 2, Supplementary Discussion for IR spectra and Supplementary References.

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yano, Y., Mitoma, N., Matsushima, K. et al. RETRACTED ARTICLE: Living annulative π-extension polymerization for graphene nanoribbon synthesis. Nature 571, 387–392 (2019). https://doi.org/10.1038/s41586-019-1331-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1331-z

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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