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.

  • Article
  • Published:

Orbital-resolved visualization of single-molecule photocurrent channels

Abstract

Given its central role in utilizing light energy, photoinduced electron transfer (PET) from an excited molecule has been widely studied1,2,3,4,5,6. However, even though microscopic photocurrent measurement methods7,8,9,10,11 have made it possible to correlate the efficiency of the process with local features, spatial resolution has been insufficient to resolve it at the molecular level. Recent work has, however, shown that single molecules can be efficiently excited and probed when combining a scanning tunnelling microscope (STM) with localized plasmon fields driven by a tunable laser12,13. Here we use that approach to directly visualize with atomic-scale resolution the photocurrent channels through the molecular orbitals of a single free-base phthalocyanine (FBPc) molecule, by detecting electrons from its first excited state tunnelling through the STM tip. We find that the direction and the spatial distribution of the photocurrent depend sensitively on the bias voltage, and detect counter-flowing photocurrent channels even at a voltage where the averaged photocurrent is near zero. Moreover, we see evidence of competition between PET and photoluminescence12, and find that we can control whether the excited molecule primarily relaxes through PET or photoluminescence by positioning the STM tip with three-dimensional, atomic precision. These observations suggest that specific photocurrent channels can be promoted or suppressed by tuning the coupling to excited-state molecular orbitals, and thus provide new perspectives for improving energy-conversion efficiencies by atomic-scale electronic and geometric engineering of molecular interfaces.

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: Atomic-scale measurement of photocurrent generation in a single molecule.
Fig. 2: Voltage dependence of photocurrent channels through a single molecule.
Fig. 3: Mechanism of photocurrent generation in an FBPc molecule.
Fig. 4: Control of the quantum efficiencies of photoelectron energy conversion.

Similar content being viewed by others

Data availability

All experimental data shown in Figs. 14, Extended Data Figs. 19 and Extended Data Tables 1, 2 can be found at Zenodo, https://doi.org/10.5281/zenodo.5553643, and are also available from the corresponding authors upon reasonable request.

Code availability

The code used to calculate the results shown in this work is available from the corresponding authors upon reasonable request.

References

  1. Tokita, Y., Shimura, J., Nakajima, H., Goto, Y. & Watanabe, Y. Mechanism of intramolecular electron transfer in the photoexcited Zn-substituted cytochrome c: theoretical and experimental perspective. J. Am. Chem. Soc. 130, 5302–5310 (2008).

    Article  CAS  PubMed  Google Scholar 

  2. Sariciftci, N. S., Smilowitz, L., Heeger, A. J. & Wudl, F. Photoinduced electron transfer from a conducting polymer to buckminsterfullerene. Science 258, 1474–1476 (1992).

    Article  ADS  CAS  PubMed  Google Scholar 

  3. Murphy, C. J. et al. Long-range photoinduced electron transfer through a DNA helix. Science 262, 1025–1029 (1993).

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Lindstrom, C. D. & Zhu, X.-Y. Photoinduced electron transfer at molecule–metal interfaces. Chem. Rev. 106, 4281–4300 (2006).

    Article  CAS  PubMed  Google Scholar 

  5. Ma, W., Ma, H., Peng, Y. Y., Tian, H. & Long, Y. T. An ultrasensitive photoelectrochemical platform for quantifying photoinduced electron-transfer properties of a single entity. Nat. Protoc. 14, 2672–2690 (2019).

    Article  CAS  PubMed  Google Scholar 

  6. Jones, A. L., Jiang, J. & Schanze, K. S. Excitation-wavelength-dependent photoinduced electron transfer in a π-conjugated diblock oligomer. J. Am. Chem. Soc. 142, 12658–12668 (2020).

    Article  CAS  PubMed  Google Scholar 

  7. O’Dea, J. R., Brown, L. M., Hoepker, N., Marohn, J. A. & Sadewasser, S. Scanning probe microscopy of solar cells: from inorganic thin films to organic photovoltaics. MRS Bull. 37, 642–650 (2012).

    Article  CAS  Google Scholar 

  8. Giridharagopal, R., Cox, P. A. & Ginger, D. S. Functional scanning probe imaging of nanostructured solar energy materials. Acc. Chem. Res. 49, 1769–1776 (2016).

    Article  CAS  PubMed  Google Scholar 

  9. Gerster, D. et al. Photocurrent of a single photosynthetic protein. Nat. Nanotechnol. 7, 673–676 (2012).

    Article  ADS  CAS  PubMed  Google Scholar 

  10. Takeuchi, O. et al. Microscopic description of the current–voltage characteristics of a bulk-heterojunction organic solar cell under illumination. Appl. Phys. Express 7, 021602 (2014).

    Article  ADS  CAS  Google Scholar 

  11. Coffey, D. C., Reid, O. G., Rodovsky, D. B., Bartholomew, G. P. & Ginger, D. S. Mapping local photocurrents in polymer/fullerene solar cells with photoconductive atomic force microscopy. Nano Lett. 7, 738–744 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  12. Imada, H. et al. Single-molecule laser nanospectroscopy with micro-electron volt energy resolution. Science 373, 95–98 (2021).

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Jaculbia, R. B. et al. Single-molecule resonance Raman effect in a plasmonic nanocavity. Nat. Nanotechnol. 15, 105–110 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  14. Wu, S. W., Ogawa, N. & Ho, W. Atomic-scale coupling of photons to single-molecule junctions. Science 312, 1362–1365 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  15. Zhu, S.-E. et al. Self-decoupled porphyrin with a tripodal anchor for molecular-scale electroluminescence. J. Am. Chem. Soc. 135, 15794–15800 (2013).

    Article  CAS  PubMed  Google Scholar 

  16. Cocker, T. L., Peller, D., Yu, P., Repp, J. & Huber, R. Tracking the ultrafast motion of a single molecule by femtosecond orbital imaging. Nature 539, 263–267 (2016).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  17. Yoshioka, K. et al. Real-space coherent manipulation of electrons in a single tunnel junction by single-cycle terahertz electric fields. Nat. Photonics 10, 762–765 (2016).

    Article  ADS  CAS  Google Scholar 

  18. Garg, M. & Kern, K. Attosecond coherent manipulation of electrons in tunneling microscopy. Science 367, 411–415 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  19. Zhang, R. et al. Chemical mapping of a single molecule by plasmon-enhanced Raman scattering. Nature 498, 82–86 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  20. Zhang, Y. et al. Visualizing coherent intermolecular dipole–dipole coupling in real space. Nature 531, 623–627 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  21. Imada, H. et al. Real-space investigation of energy transfer in heterogeneous molecular dimers. Nature 538, 364–367 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Doppagne, B. et al. Vibronic spectroscopy with submolecular resolution from STM-induced electroluminescence. Phys. Rev. Lett. 118, 127401 (2017).

    Article  ADS  PubMed  Google Scholar 

  23. Kimura, K. et al. Selective triplet exciton formation in a single molecule. Nature 570, 210–213 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  24. Imada, H. et al. Single-molecule investigation of energy dynamics in a coupled plasmon–exciton system. Phys. Rev. Lett. 119, 013901 (2017).

    Article  ADS  PubMed  Google Scholar 

  25. Murray, C. et al. Infrared and Raman spectroscopy of free-base and zinc phthalocyanines isolated in matrices. Phys. Chem. Chem. Phys. 12, 10406–10422 (2010).

    Article  CAS  PubMed  Google Scholar 

  26. Murray, C. et al. Visible luminescence spectroscopy of free-base and zinc phthalocyanines isolated in cryogenic matrices. Phys. Chem. Chem. Phys. 13, 17543–17554 (2011).

    Article  CAS  PubMed  Google Scholar 

  27. Imai-Imada, M. et al. Energy-level alignment of a single molecule on ultrathin insulating film. Phys. Rev. B 98, 201403 (2018).

    Article  ADS  CAS  Google Scholar 

  28. Doppagne, B. et al. Electrofluorochromism at the single-molecule level. Science 361, 251–255 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  29. Repp, J., Meyer, G., Stojković, S. M., Gourdon, A. & Joachim, C. Molecules on insulating films: scanning-tunneling microscopy imaging of individual molecular orbitals. Phys. Rev. Lett. 94, 026803 (2005).

    Article  ADS  PubMed  CAS  Google Scholar 

  30. Ikeda, T., Iino, R. & Noji, H. Real-time fluorescence visualization of slow tautomerization of single free-base phthalocyanines under ambient conditions. Chem. Commun. 50, 9443–9446 (2014).

    Article  CAS  Google Scholar 

  31. Liljeroth, P., Repp, J. & Meyer, G. Current-induced hydrogen tautomerization and conductance switching of naphthalocyanine molecules. Science 317, 1203–1206 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  32. Doppagne, B. et al. Single-molecule tautomerization tracking through space- and time-resolved fluorescence spectroscopy. Nat. Nanotechnol. 15, 207–211 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  33. Böckmann, H. et al. Direct observation of photoinduced tautomerization in single molecules at a metal surface. Nano Lett. 16, 1034–1041 (2016).

    Article  ADS  CAS  PubMed  Google Scholar 

  34. Miwa, K., Najarian, A. M., Mccreery, R. L. & Galperin, M. Hubbard nonequilibrium Green’s function analysis of photocurrent in nitroazobenzene molecular junction. J. Phys. Chem. Lett. 10, 1550–1557 (2019).

    Article  CAS  PubMed  Google Scholar 

  35. Miwa, K. et al. Many-body state description of single-molecule electroluminescence driven by a scanning tunneling microscope. Nano Lett. 19, 2803–2811 (2019).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Yang, B. et al. Sub-nanometre resolution in single-molecule photoluminescence imaging. Nat. Photonics 14, 693–699 (2020).

    Article  ADS  CAS  Google Scholar 

  37. Qiu, X. H., Nazin, G. V. & Ho, W. Vibrationally resolved fluorescence excited with submolecular precision. Science 299, 542–546 (2003).

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Kuhnke, K., Große, C., Merino, P. & Kern, K. Atomic-scale imaging and spectroscopy of electroluminescence at molecular interfaces. Chem. Rev. 117, 5174–5222 (2017).

    Article  CAS  PubMed  Google Scholar 

  39. Yang, B., Kazuma, E., Yokota, Y. & Kim, Y. Fabrication of sharp gold tips by three-electrode electrochemical etching with high controllability and reproducibility. J. Phys. Chem. C 122, 16950–16955 (2018).

    Article  CAS  Google Scholar 

  40. Miwa, K., Imada, H., Kawahara, S. & Kim, Y. Effects of molecule–insulator interaction on geometric property of a single phthalocyanine molecule adsorbed on an ultrathin NaCl film. Phys. Rev. B 93, 165419 (2016).

    Article  ADS  CAS  Google Scholar 

  41. Neuman, T., Esteban, R., Casanova, D., García-Vidal, F. J. & Aizpurua, J. Coupling of molecular emitters and plasmonic cavities beyond the point-dipole approximation. Nano Lett. 18, 2358–2364 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  42. Frisch, M. J. et al. Gaussian 16, revision C.01 (Gaussian, Inc., 2016); https://gaussian.com.

  43. Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 90, 1007–1023 (1989).

    Article  ADS  CAS  Google Scholar 

  44. Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 98, 5648–5652 (1993).

    Article  ADS  CAS  Google Scholar 

  45. Henderson, T. M., Izmaylov, A. F., Scalmani, G. & Scuseria, G. E. Can short-range hybrids describe long-range-dependent properties? J. Chem. Phys. 131, 044108 (2009).

    Article  ADS  CAS  PubMed  Google Scholar 

  46. Baer, R., Livshits, E. & Salzner, U. Tuned range-separated hybrids in density functional theory. Annu. Rev. Phys. Chem. 61, 85–109 (2010).

    Article  CAS  PubMed  Google Scholar 

  47. Runge, E. & Gross, E. K. U. Density-functional theory for time-dependent systems. Phys. Rev. Lett. 52, 997–1000 (1984).

    Article  ADS  CAS  Google Scholar 

  48. Casida, M. E. Time-dependent density functional response theory for molecules. In Recent Advances in Density Functional Methods: Part I (ed. Chong, D. P.) 155–192 (World Scientific, 1995).

  49. Hirata, S. & Head-Gordon, M. Time-dependent density functional theory within the Tamm–Dancoff approximation. Chem. Phys. Lett. 314, 291–299 (1999).

    Article  ADS  CAS  Google Scholar 

  50. Santoro, F., Improta, R., Lami, A., Bloino, J. & Barone, V. Effective method to compute Franck–Condon integrals for optical spectra of large molecules in solution. J. Chem. Phys. 126, 084509 (2007).

    Article  ADS  PubMed  CAS  Google Scholar 

  51. Santoro, F., Lami, A., Improta, R. & Barone, V. Effective method to compute vibrationally resolved optical spectra of large molecules at finite temperature in the gas phase and in solution. J. Chem. Phys. 126, 184102 (2007).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Santoro, F., Lami, A., Improta, R., Bloino, J. & Barone, V. Effective method for the computation of optical spectra of large molecules at finite temperature including the Duschinsky and Herzberg–Teller effect: the Qx band of porphyrin as a case study. J. Chem. Phys. 128, 224311 (2008).

    Article  ADS  CAS  PubMed  Google Scholar 

  53. Barone, V., Bloino, J., Biczysko, M. & Santoro, F. Fully integrated approach to compute vibrationally resolved optical spectra: from small molecules to macrosystems. J. Chem. Theory Comput. 5, 540–554 (2009).

    Article  CAS  PubMed  Google Scholar 

  54. Scivetti, I. & Persson, M. Frontier molecular orbitals of a single molecule adsorbed on thin insulating films supported by a metal substrate: electron and hole attachment energies. J. Phys. Condens. Matter 29, 355002 (2017).

    Article  PubMed  Google Scholar 

  55. Galperin, M. Photonics and spectroscopy in nanojunctions: a theoretical insight. Chem. Soc. Rev. 46, 4000–4019 (2017).

    Article  CAS  PubMed  Google Scholar 

  56. Miwa, K., Chen, F. & Galperin, M. Towards noise simulation in interacting nonequilibrium systems strongly coupled to baths. Sci. Rep. 7, 9735 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  57. Chen, F., Ochoa, M. A. & Galperin, M. Nonequilibrium diagrammatic technique for Hubbard Green functions. J. Chem. Phys. 146, 92301 (2017).

    Article  CAS  Google Scholar 

  58. Cohen, G. & Galperin, M. Green’s function methods for single molecule junctions. J. Chem. Phys. 152, 090901 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  59. White, A. J., Ochoa, M. A. & Galperin, M. Nonequilibrium atomic limit for transport and optical response of molecular junctions. J. Phys. Chem. C 118, 11159–11173 (2014).

    Article  CAS  Google Scholar 

  60. Schulz, F. et al. Many-body transitions in a single molecule visualized by scanning tunnelling microscopy. Nat. Phys. 11, 229–234 (2015).

    Article  CAS  Google Scholar 

  61. Ervasti, M. M., Schulz, F., Liljeroth, P. & Harju, A. Single- and many-particle description of scanning tunneling spectroscopy. J. Electron Spectros. Relat. Phenom. 219, 63–71 (2017).

    Article  CAS  Google Scholar 

  62. Seldenthuis, J. S., van der Zant, H. S. J., Ratner, M. A. & Thijssen, J. M. Electroluminescence spectra in weakly coupled single-molecule junctions. Phys. Rev. B 81, 205430 (2010).

    Article  ADS  CAS  Google Scholar 

  63. Fatayer, S. et al. Reorganization energy upon charging a single molecule on an insulator measured by atomic force microscopy. Nat. Nanotechnol. 13, 376–380 (2018).

    Article  ADS  CAS  PubMed  Google Scholar 

  64. Yu, P., Kocić, N., Repp, J., Siegert, B. & Donarini, A. Apparent reversal of molecular orbitals reveals entanglement. Phys. Rev. Lett. 119, 56801 (2017).

    Article  ADS  Google Scholar 

  65. Wu, S. W., Nazin, G. V., Chen, X., Qiu, X. H. & Ho, W. Control of relative tunneling rates in single molecule bipolar electron transport. Phys. Rev. Lett. 93, 236802 (2004).

    Article  ADS  CAS  PubMed  Google Scholar 

  66. Novotny, L. A. & Hecht, B. Principles of Nano-Optics (Cambridge Univ. Press, 2012).

Download references

Acknowledgements

We thank B. Yang, E. Kazuma, Y. Hasegawa and Y. Shimizu for supporting the preparation of the Au tips. Some of the numerical calculations were carried out using the HOKUSAI system at RIKEN. All experiments were conducted at RIKEN in Japan. This work was financially supported in part by JST PRESTO (grant no. JPMJPR186225 (H.I.)) and by Japan Society for the Promotion of Science (JSPS) KAKENHI: grant nos JP 19K23591 (M.I.-I.), JP19H04681, JP 17H04796 (H.I.), JP 20K22488 (K.K.), JP 18K14153 (R.B.J.), JP 18H05257 (Y.K.), JP17H06173, JP17H05430 (M.U.), JP20H02728 (A.M.).

Author information

Authors and Affiliations

Authors

Contributions

M.I.-I, H.I. and Y.K. conceived the project and constructed the experimental set-up with the help of Y.K. M.I.-I., H.I., K.K., I.Z., R.B.J. and H.Y. performed the experiment using STM and analysed the data. Y.T., A.M. and M.U. synthesized and characterized the D2Pc molecule. K.M. conducted the theoretical calculations. Y.K. directed the project. All authors discussed the results and wrote the manuscript.

Corresponding authors

Correspondence to Hiroshi Imada or Yousoo Kim.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Leo Gross, Christoph Lienau and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Trace of tip–height variation under laser irradiation.

The laser power was 80 μW, and the energy was 1,816 meV. During the measurement, the tip was placed on the Ag(111) with the closed STM feedback loop.

Extended Data Fig. 2 Voltage dependence of dark-current images over a single molecule.

Current images over the FBPc molecule at −2.0 and 0.0 V (image size, 3.0 nm × 3.0 nm). While the tip was scanning, the laser was turned off, and the STM feedback loop was open. The tip–molecule distance was maintained to be the same as the laser-on image in Fig. 2b. Shades of blue and red indicate current flow from the substrate to the tip and vice versa, respectively. The line profiles along the dashed lines are shown below the dark-current images. The red dots show the raw data at each pixel and the black line shows a smoothed profile by a moving average of five adjacent data points. At −2.0 V, the dark current slightly exceeds the margin of error around the centre of the molecule, whereas it does not at 0.0 V. The maximum absolute value of the detected current at −2.0 V was about 0.03 pA. Because this value was only 0.3% of the maximum detected current value under illumination (Fig. 2b), the contribution of the dark current to the photocurrent image in Fig. 2b is negligibly small.

Extended Data Fig. 3 Tautomerization reaction of an FBPc molecule under laser irradiation.

a, Z(t) spectra were measured on a D2Pc molecule on the NaCl(4ML)/Ag(111) with the STM feedback loop closed (Vs = +1.0 V, It = 3 pA). The black and blue spectra were measured under laser-off and laser-on conditions, respectively. The incident laser energy was tuned at the resonance of the S0–S1 transition shown in Fig. 1d, with the power at 1 μW. Both spectra were measured with the same tip placed at the same distance and angle from the centre of the molecule, and the Z values for both spectra refer to the identical position. b, The current images of a D2Pc molecule (Vs = +0.75 V) under laser-on (left) and laser-off (right) conditions, respectively. The power of the laser was 2 μW. c, The orientation of the molecule in the images shown in b. Under the laser-on condition, tautomerization reactions cause the molecular axes to rotate through 90°, whereas under laser-off condition it does not.

Extended Data Fig. 4 Molecular structure and the spatial distribution of frontier molecular orbitals.

Molecular structure of FBPc and the spatial distributions of the HOMO, LUMO and nearly degenerated LUMO+1 of FBPc in the gas phase as predicted by density functional theory (DFT) calculations. The orientation of molecule and the position of the arrows, which show the position 45° from the axes, correspond to those in Fig. 2b, d.

Extended Data Fig. 5 Voltage dependence of the photocurrent image in the transition range.

Photocurrent images over the FBPc molecule at −0.4, −0.3, −0.25, −0.2 and −0.1 V (Image size: 2.4 nm × 2.4 nm) acquired with the STM feedback loop open. The laser energy was set at the resonance of the S0–S1 transition, with a power of 77 μW. Shades of blue and red represent the current flow from the substrate to the tip and vice versa, respectively. The line profiles along the dashed line in each photocurrent image are shown below the respective images. The vertical dashed lines in the line profiles are for comparing the same tip positions. It is revealed that positions showing the local maximum of the positive photocurrent at −0.1 V correspond to where the negative photocurrent value shows a local minimum at −0.4 V. This result is indicative of the competition between the negative and positive photocurrent channels in the molecule.

Extended Data Fig. 6 Line profiles of photocurrent and photoluminescence signals.

a, Photocurrent image of an FBPc molecule on NaCl(4ML)/Ag(111) measured at Vs = −2.0 V and the STM tip position during the measurement in Fig. 4d, e (black circles). b, The line profile of the photocurrent along the dashed line in a (45° from the molecular axes). Photocurrent signals were measured at lateral tip positions ranging from 0.6 nm to 2.4 nm. The size of the photocurrent signal agrees well with the molecular size observed in the STM image, indicating that photocurrent generation occurs when the tip is placed on the molecule. c, The line profile of photoluminescence (PL) along the dashed line in a. Photoluminescence signals appear at all lateral tip positions between 0.0 nm and 3.0 nm. The intensity is minimal at the centre of the molecule (~1.5 nm), and increases with distance from the centre. The local maximum values appear at ~0.5 and ~2.6 nm, and the intensity further turns to decrease with distance. These tendencies correspond to the tip–position dependence of the coupling between the molecule and the localized plasmon which drives the molecular luminescence36. Because the photoluminescence is driven by the localized plasmon, which is several nanometres long24, photoluminescence can be observed even when the tip is outside the molecule. The STM tip positions during the measurement in Fig. 4d, e are shown by the arrows. Although both photocurrent and photoluminescence signals were observed on the molecule (at the position of the left arrow), only the photoluminescence signal was detected outside the molecule (at the right arrow). Using this difference, we investigated the influence of photocurrent generation on photoluminescence quenching (Fig. 4d, e).

Extended Data Fig. 7 Comparison of the populations of increasing electrons and decreasing photons.

The increase of electrons (red area in Fig. 4d) and the decrease of photons (blue area in Fig. 4d) are plotted against Ztip–mol. The number of electrons was obtained from the detected photocurrent value divided by the elementary charge e. The number of quenched photons (\({I}_{{\rm{ph}}}^{{\rm{quench}}}\)) was obtained by the following equation. \({I}_{{\rm{ph}}}^{{\rm{quench}}}=({I}_{{\rm{ph}}}^{{\rm{\det }}}-\,{I}_{{\rm{ph}}}^{{\rm{fit}}})/{\eta }_{{\rm{\det }}}\). Here, \({I}_{{\rm{ph}}}^{{\rm{\det }}}\) is the number of the photons (counts per second) detected in the experiment, and \({I}_{{\rm{ph}}}^{{\rm{fit}}}\) is the photon numbers (counts per second) deduced by the fitting curve Iph = 1.09 × 105 × exp(−6.77Ztip–mol) for the Ztip–mol range between 1.1 nm and 0.55 nm shown in Fig. 4d. By extrapolating the fitted curve into the Ztip–mol value less than 0.53 nm, we estimated the expected photon intensity without the photoluminescence quenching in this region. ηdet is the detection efficiency determined by the experimental set-up. By considering the collection solid angle of the lens, detection quantum efficiency of the detector, and the reflection, diffraction and transmission of the optics, ηdet was estimated to be 4.9 × 10−4 for this measurement (see Methods). It was revealed that the number of electrons flowing as photocurrent and the number of quenched photons at each Ztip–mol were comparable. These results suggest that the observed photoluminescence quenching mainly originated from the photocurrent generation. The slightly large value of the quenched photons might come from another nonradiative recombination pathway or underestimation of ηdet.

Extended Data Fig. 8 Determination of the tip–molecule distance.

a, The conductance (G) curve for the bare Ag(111) (black) and the linear fitting curve (red). The tip displacement Z refers to the initial tip position at Vs = −2.5 V and It = 3 pA. The G is normalized by the quantization conductance G0 which is defined by 2e2/h with the electron charge e and Planck’s constant h. b, A schematic illustration of the various relevant distances in the photocurrent measurement.

Extended Data Fig. 9 Simulation results of the IV curve in the single-molecule junction under laser irradiation.

af, Results are given for various values of electron–phonon coupling parameter λ.

Extended Data Table 1 Total energy Etotal (summation of the electronic and zero-point vibrational energies) for each electronic state
Extended Data Table 2 The correction \({{\boldsymbol{E}}}_{{\boldsymbol{N}}{\boldsymbol{a}}}^{{\bf{Img}}}\) of the total energy for each molecular many-body state owing to the image-charge effects from the electrodes

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Imai-Imada, M., Imada, H., Miwa, K. et al. Orbital-resolved visualization of single-molecule photocurrent channels. Nature 603, 829–834 (2022). https://doi.org/10.1038/s41586-022-04401-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-04401-0

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