A switchable system between magnetic and natural circularly polarised luminescence via J-aggregation using photosynthetic antenna model compounds

Toranosuke Tomikawa a, Yuichi Kitagawa *b, Koki Yoshioka d, Kei Murata c, Tomohiro Miyatake d, Yasuchika Hasegawa *be and Kazuyuki Ishii *c
aGraduate School of Chemical Sciences and Engineering, Hokkaido University, Kita 13, Nishi 8, Kita-ku, Sapporo, Hokkaido 060-8628, Japan
bFaculty of Engineering, Hokkaido University, Kita 13, Nishi 8, Kita-ku, Sapporo, Hokkaido 060-8628, Japan
cInstitute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan. E-mail: k-ishii@iis.u-tokyo.ac.jp
dDepartment of Materials Chemistry, Ryukoku University, 1-5 Yokotani, Seta Oecho, Otsu, Shiga 520-2194, Japan
eInstitute for Chemical Reaction Design and Discovery (WPI-ICReDD), Hokkaido University, Kita 21, Nishi 10, Kita-ku, Sapporo, Hokkaido 001-0021, Japan

Received 14th November 2022 , Accepted 13th January 2023

First published on 25th January 2023


Abstract

A switchable system between magnetic and natural circularly polarized luminescence via J-aggregation using photosynthetic antenna model compounds is demonstrated for the first time. The J-aggregation provides characteristic dispersion-type circularly polarised luminescence signals based on the degenerate J-band (gCPL = 0.03).


Chiral compounds exhibit circularly polarised luminescence (CPL), which is characterized by the differential emission of right- and left-handed circularly polarised light.1,2 CPL has attracted considerable attention because of its applications in security tags, sensors, and organic electroluminescent devices for 3D displays.3–9

The magnitude of CPL can be expressed using the dissymmetry factor (gCPL).1,2 The gCPL value depends on the transition electric dipole moment ([small mu, Greek, vector]), transition magnetic dipole moment ([m with combining right harpoon above (vector)]), and angle between [small mu, Greek, vector] and [m with combining right harpoon above (vector)]. Although the combination of enhanced |[small mu, Greek, vector]| and |[m with combining right harpoon above (vector)]| offers a large gCPL value with bright emission, achieving this is still challenging in modern chemistry because the linear electronic motion of [small mu, Greek, vector] is fundamentally different from the rotational electronic motion of [m with combining right harpoon above (vector)]. Chiral J-type aggregation is a promising strategy for achieving large |[small mu, Greek, vector]| and |[m with combining right harpoon above (vector)]| values based on the exciton chirality theory.10–17 Ghosh et al. recently reported a large gCPL (= 4.6 × 10−2) with a high emission quantum yield (Φ = 43%) using helical J-aggregates composed of chiral naphthalene-diimide derivatives.10 Moreover, the flexibility of aggregation can introduce switching functions. For example, Harada et al. have achieved flexible CPL switching behaviour using pseudo cyanine J-aggregates by thermal stimulus.13 Using thiopheneboronic acid, Kameta et al. demonstrated thermal stimulus-driven CPL switching via morphological transformation from H- to J-aggregates (gCPL ≈ 0 → gCPL = 3.1 × 10−3).14 Thus, flexible CPL switching systems with a large gCPL value can open up new possibilities for photonic applications.

Magnetic CPL (MCPL) is the CPL phenomenon under a magnetic field. MCPL is observed in both chiral and achiral compounds,18–20 and the magnitude of MCPL for organic compounds is related to the orbital angular momentum of the aromatic π-electron system.21,22 Specifically, the multi-function switching based on both natural CPL and MCPL is expected to provide new insights for the development of magneto-photonic materials based on polarised light.23–25

Herein, we provide a novel system, in which MCPL and natural CPL are switched upon J-aggregation. To demonstrate our MCPL/CPL switching concept, we employed chiral Zn-chlorin (zinc 3-devinyl-3-methoxymethyl pyropheophorbide-a monoester: ZnChl, Fig. 1a) with a large aromatic π-electron system because of the following reasons.26 (1) The interaction of Zn(II) with a methoxy-substituent gives the formation of J-aggregates (Fig. 1b), which correspond to a model compound of light-harvesting antennas in green photosynthetic bacteria, i.e., chlorosomes.27 (2) The transition to the lowest excited state in the chlorin unit, called the Qy band, has a large transition probability,28 resulting in strong light absorption and emission properties as well as good chiroptical properties in the Qy band via J-aggregation.29 (3) The large orbital angular momentum derived from the large aromatic π-electron system of the chlorin unit imparts effective magneto-optical properties to the Qy band, which is weakened by the J-aggregation.29 Based on these facts, the J-aggregation of ZnChl should be an appropriate system for realizing the switch between MCPL and natural CPL (Fig. 1c).


image file: d2tc04841h-f1.tif
Fig. 1 (a) Molecular structure of ZnChl and (b) schematic image of the chiral J-aggregates of ZnChls.31–38 (c) Magnetic and natural circularly polarised luminescence switching via J-aggregation (Red arrow: transition electric dipole moment).

ZnChl was prepared according to the previous method.26 Here, the Zn ion was employed as the central metal ion instead of a Mg ion, since Zn chlorins are more stable than Mg chlorins.30 The J-aggregates of ZnChls were prepared by diluting the methanol solution of ZnChl with 99-fold volume of water (ESI).31–38 The electronic absorption, emission, MCPL, and CPL spectra of the ZnChl monomer are shown in Fig. 2. The absorption bands at 581, 611, and 658 nm were assigned to the Qx(0,0), Qy(1,0), and Qy(0,0) bands, respectively (Fig. 2a).29 Strong emission bands were observed at 665 and 720 nm, corresponding to their Qy bands (Fig. 2b). The emission quantum yield and lifetime were 0.067 and 4.7 ns, respectively. The Qy emission bands showed a distinguishable, positive MCPL signal (Fig. 2c, gCPL = 0.003) in contrast to the negligible CPL signal (Fig. 2c) derived from the poor chiroptical property of the asymmetric centre at positions 17 and 18 (Fig. 1a).


image file: d2tc04841h-f2.tif
Fig. 2 (a) Electronic absorption spectra of the ZnChl monomer (dotted line) and J-aggregates of ZnChls (solid line). (b) Emission spectra of the ZnChl monomer (dotted line, λex = 430 nm) and J-aggregates of ZnChls (solid line, λex = 450 nm). (c) CPL (blue line) and MCPL (red line) spectra of the ZnChl monomer (λex = 430 nm). (d) CPL (blue line) and MCPL (red line) spectra of J-aggregates of ZnChls (λex = 450 nm).

In our previous study,29 the electronic transition bands were calculated using the time-dependent density functional theory (TD-DFT, B3LYP/6-31G(d)) method. In the calculation, we employed the structure of a ZnChl derivative having an ethyl group at position 17. The spectroscopic properties of the Q bands can be reasonably explained by Gouterman's model based on four orbitals (b1, b2, c2, and c1) and four electronic configurations ((b1c2), (b2c2), (b1c1), and (b2c1)), as shown in Fig. 3a.39,40 Because the energy difference between the (b2c2) and (b1c1) configurations is relatively large, the Qy transition mainly originates from the single electronic (b2c2) configuration (Table S1 in ref. 29). On the other hand, the (b1c2) configuration is heavily admixed with the (b2c1) configuration, resulting in weak Qx bands. Thus, with reference to the magnetic circular dichroism intensity, a strong Faraday B-type MCPL signal is observed for the Qy emission band because of the magnetic moment between the Qx and Qy bands; this magnetic moment originates from the large orbital angular momentum of the aromatic π-electron system (i.e., 〈b2|lz|b1〉 and 〈c2|lz|c1〉).22


image file: d2tc04841h-f3.tif
Fig. 3 (a) Molecular orbitals of ZnChl.29 (b) Energy diagram of the ZnChl monomer and their J-aggregates for explaining the ineffective MCPL due to J-aggregation. The molecular orbitals (isosurface value = 0.02) were calculated by the DFT method (B3LYP/6-31G(d)).

In the electronic absorption spectrum of the chiral J-aggregates of ZnChls, the Qy band remarkably shifted to the red-side (729 nm) owing to the exciton interaction between the ZnChl units, whereas the shift of the Qx band was very small. The narrow J-band (FWHM of 280 cm−1) reflects the formation of well-ordered J-aggregates. Consequently, the J-aggregation promotes the energy separation between the Qy and Qx bands (monomer, 2120 cm−1; J-aggregates, 3510 cm−1). In the emission spectrum, a sharp signal was observed at 732 nm with a small Stokes shift (ΔE = 60 cm-1), as shown in Fig. 2b. The emission quantum yield was 0.004.41 In the emissive J-band region, no distinguishable MCPL signal was observed, although the ZnChl monomer exhibited MCPL (Fig. 2c). This is reasonably explained by the increase in the energy gap between the Qx and Qy bands via J-aggregation, because the MCPL B term, which is proportional to Im((〈Qy|m|Qx〉/(E(Qx) − E(Qy)))〈S0|μ|Qy〉〈Qx|μ|S0〉), is in inverse proportion to the energy gap, E(Qx) − E(Qy) (〈Qy|m|Qx〉: the magnetic moment between the Qx and Qy bands, 〈S0|μ|Qi〉: the electric dipole moment of the Qi band, E(Qi): the Qi band energy). On the other hand, an opposite trend was observed for CPL. Whereas the ZnChl monomer exhibited negligible CPL signal, the chiral J-aggregates of ZnChls exhibited intense positive/negative CPL signals in the J-band region (Fig. 2d, gCPL = 0.03, positive peak: 723.5 nm, negative peak: 739 nm). The CPL spectrum is similar to the dispersion-type CD spectral pattern in the J-band region (Fig. S1, positive peak: 728.5 nm, negative peak: 735 nm, ESI), which indicates the coexistence of the lowest excited singlet state and the nearby excited singlet state in the J-band region. The energy splitting (120 cm−1) between these excited states is comparable to the thermal energy (∼200 cm−1), and therefore, the dispersion-type CPL spectral pattern was observed as a sum of the negative CPL from the lowest excited singlet state and the positive CPL from the thermally activated, nearby excited singlet state. Thus, using photosynthetic antenna model compounds, we successfully demonstrate the first system in which the MCPL/CPL behaviour is switched upon J-aggregation (Fig. 2c and d).

Finally, we attempted to theoretically clarify the difference in gCPL between the monomer and J-aggregates. gCPL is given by

 
image file: d2tc04841h-t1.tif(1)
where IL and IR are the intensities of the left- and right-handed circularly polarised light, respectively. The ZnChl monomer did not show a distinguishable CPL signal because of its quite small |[m with combining right harpoon above (vector)]|. In the case of J-aggregates, the gCPL value is re-written using the following equations:
 
image file: d2tc04841h-t2.tif(2)
 
image file: d2tc04841h-t3.tif(3)
 
image file: d2tc04841h-t4.tif(4)

Here, [small mu, Greek, vector]J and [m with combining right harpoon above (vector)]J are the transition electric dipole moment and transition magnetic dipole moment of the J-aggregates, respectively, and the label (p) denotes the p-th ZnChl unit in the J-aggregates (Fig. 4a). According to the exciton chirality method,12 the [m with combining right harpoon above (vector)]J is expressed using the following equation:

 
image file: d2tc04841h-t5.tif(5)
Here, [r with combining right harpoon above (vector)]0(p) and ωna(p) are the position vector and angular frequency of the transition energy for the p-th unit, respectively. Thus, the |[m with combining right harpoon above (vector)]J| value can be approximated by the |[small mu, Greek, vector](p)| values, which means solving the dilemma between the linear electronic motion of [small mu, Greek, vector] and the rotational electronic motion of [m with combining right harpoon above (vector)]. To obtain information about the configurations between the ZnChl units, geometry optimisation was calculated for the ZnChl dimer. In the optimisation process, molecular mechanics (UFF) was initially used to find the energetic minimum of the dimer. Then, the geometry optimisation was calculated using the Hartree–Fock method (3-21G), followed by the DFT method (B3LYP/6-31G+(d,p)). In the optimised structure, the aromatic ZnChl planes were parallel to each other, but the in-plane axes were twisted: the twisted angle was estimated to be ∼43° (Fig. 4b). Based on these equations and the structural configuration, the large gCPL value can be reasonably explained by the twisted configuration between the ZnChl units with a large |[small mu, Greek, vector]| value.


image file: d2tc04841h-f4.tif
Fig. 4 (a) Schematic of ZnChl stacking (Red arrow: transition electric dipole moment). (b) The optimised structure of a ZnChl dimer. In the optimisation process, molecular mechanics (UFF) was initially used to find the energetic minimum of the dimer. Then, the geometry optimisation was calculated using the Hartree–Fock method (3-21G), followed by the DFT method (B3LYP/6-31+G(d,p)).

Conclusions

Using ZnChl, the switching of MCPL/CPL via J-aggregation has been clearly demonstrated for the first time. The effective exciton coupling between the ZnChl units triggers the novel switching property. These results provide new insights into the design of CPL switching materials even in solid media, and are expected to promote the development of a novel research area based on the science and technology of polarised-light.

Author contributions

K. I. designed research. K. Y. and K. M. prepared samples. T. T., Y.K., T. M., and Y. H. performed optical measurements. T. T., Y. K., K. M., T. M., Y. H., and K. I. wrote the paper. All authors reviewed the paper.

Conflicts of interest

There are no conflicts of interest to declare.

Acknowledgements

This work was supported by the Institute for Chemical Reaction Design and Discovery (ICReDD), established by the World Premier International Research Center Initiative (WPI). This work was also supported by the special fund of Institute of Industrial Science, The University of Tokyo and the special fund of Research Centre for Interdisciplinary Studies in Religion, Science and Humanities, Ryukoku University.

Notes and references

  1. F. S. Richardson and J. P. Riehl, Chem. Rev., 1977, 77, 773 CrossRef CAS.
  2. J. P. Riehl and F. S. Richardson, Chem. Rev., 1986, 86, 1 CrossRef CAS.
  3. Y. Kitagawa, S. Wada, M. D. J. Islam, K. Saita, M. Gon, K. Fushimi, K. Tanaka, S. Maeda and Y. Hasegawa, Commun. Chem., 2020, 3, 119 CrossRef CAS.
  4. F. Zhang, Q. Li, C. Wang, D. Wang, M. Song, Z. Li, X. Xue, G. Zhang and G. Qing, Adv. Funct. Mater., 2022, 32, 2204487 CrossRef CAS.
  5. G. Muller, Dalton Trans., 2009, 9692 RSC.
  6. Y. Imai, Y. Nakano, T. Kawai and J. Yuasa, Angew. Chem., Int. Ed., 2018, 57, 8973 CrossRef CAS PubMed.
  7. F. Zinna, M. Pasini, F. Galeotti, C. Botta, L. D. Bari and U. Giovanella, Adv. Funct. Mater., 2017, 27, 1603719 CrossRef.
  8. F. Song, Z. Xu, Q. Zhang, Z. Zhao, H. Zhang, W. Zhao, Z. Qiu, C. Qi, H. Zhang, H. H. Y. Sung, I. D. Williams, J. W. Y. Lam, Z. Zhao, A. Qin, D. Ma and B. Z. Tang, Adv. Funct. Mater., 2018, 28, 1800051 CrossRef.
  9. Y. Sang, J. Han, T. Zhao, P. Duan and M. Liu, Adv. Mater., 2020, 32, 1900110 CrossRef CAS PubMed.
  10. A. Mukherjee and S. Ghosh, Chem. – Eur. J., 2020, 26, 1287 Search PubMed.
  11. Y. Kitagawa, H. Segawa and K. Ishii, Angew. Chem., Int. Ed., 2011, 50, 9133 CrossRef CAS PubMed.
  12. N. Harada and N. Koji, Circular dichroic spectroscopy: exciton coupling in organic stereochemistry, Tokyo Kagaku Dojin, Tokyo, 1982 Search PubMed.
  13. T. Harada, M. Kurihara, R. Kuroda and H. Moriyama, Chem. Lett., 2012, 41, 1442 CrossRef CAS.
  14. N. Kameta and T. Shimizu, Nanoscale, 2020, 12, 2999 RSC.
  15. M. Morita, D. Rau, H. Takeba and M. Herren, Microelectron. Eng., 2000, 51–52, 605 CrossRef CAS.
  16. M. Gon, R. Sawada, Y. Morisaki and Y. Chujo, Macromolecules, 2017, 50, 1790 CrossRef CAS.
  17. T. Yamada, K. Nomura and M. Fujiki, Macromolecules, 2018, 51, 2377 CrossRef CAS.
  18. R. A. Shatwell and A. J. McCaffery, J. Chem. Soc., Chem. Commun., 1973, 15, 546 RSC.
  19. J. P. Riehl and F. S. Richardson, J. Chem. Phys., 1976, 65, 1011 CrossRef CAS.
  20. J. P. Riehl and F. S. Richardson, J. Chem. Phys., 1977, 66, 1988 CrossRef CAS.
  21. D. H. Metcalf, T. C. VanCott, S. W. Synder, P. N. Schatz and B. E. Williamson, J. Phys. Chem., 1990, 94, 2828 CrossRef CAS.
  22. J. Mack, M. J. Stillman and N. Kobayashi, Coord. Chem. Rev., 2007, 251, 429 CrossRef CAS.
  23. S. Pucher, C. Liedl, S. Jin, A. Rauschenbeutel and P. Schneeweiss, Nat. Photonics, 2022, 16, 380 CrossRef CAS.
  24. N. B. Vilas, C. Hallas, L. Anderegg, P. Robichaud, A. Winnicki, D. Mitra and J. M. Doyle, Nature, 2022, 606, 70 CrossRef CAS PubMed.
  25. P. Zhang, T. F. Chung, Q. Li, S. Wang, Q. Wang, W. L. B. Huey, S. Yang, J. Goldberger, J. Yao and X. Zhang, Nat. Mater., 2022, 21, 1373–1378 CrossRef CAS PubMed.
  26. T. Miyatake, S. Tanigawa, S. Kato and H. Tamiaki, Tetrahedron Lett., 2007, 48, 2251 CrossRef CAS.
  27. T. Miyatake and H. Tamiaki, Coord. Chem. Rev., 2010, 254, 2593 CrossRef CAS.
  28. S. Ogasawara, K. Nakano and H. Tamiaki, Eur. J. Org. Chem., 2020, 5537 CrossRef CAS.
  29. Y. Kitagawa, T. Miyatake and K. Ishii, Chem. Commun., 2012, 48, 5091 RSC.
  30. S. Patwardhan, S. Sengupta, F. Würthner, L. D. A. Siebbeles and F. Grozema, J. Phys. Chem. C, 2010, 114, 20834 CrossRef CAS.
  31. Chlorophyll self-aggregates in the light-harvesting system of photosynthetic bacteria have been extensively studied, but their precise supramolecular structure has not been clarified by X-ray crystallography.32 Balaban performed single-crystal X-ray crystallographic studies of a porphyrin-type chlorophyll model compound, which strongly supported the parallel-type supramolecular structure linked by intermolecular coordination bonds.33 Tamiaki et al. reported that chlorophyll pigments without the coordinative peripheral substituent and/or central metal did not afford chlorophyll aggregates.34 In addition, molecular modeling studies35,36 and NMR correlation experiments37 have supported the supramolecular structure. Thus, many studies on chlorophyllous aggregates have been discussed based on this structural model.38 In the amphiphilic chlorophyll derivative used in this study, the presence of a coordinative methoxy group and central zinc is essential for forming the self-aggregates.26 Consequently, the supramolecular structure shown in Fig. 1 of our manuscript is based on the widely supported structural model of the chlorophyll aggregate.
  32. S. Matsubara and H. Tamiaki, J. Photochem. Photobiol., C, 2020, 45, 100385 CrossRef CAS.
  33. T. S. Balaban, Acc. Chem. Res., 2005, 38, 612 CrossRef CAS PubMed.
  34. H. Tamiaki, M. Amakawa, Y. Shimono, R. Tanikaga, A. R. Holzwarth and K. Schaffner, Photochem. Photobiol., 1996, 63, 92 CrossRef CAS.
  35. A. Holzwarth and K. Schaffner, Photosynth. Res., 1994, 41, 225 CrossRef CAS PubMed.
  36. J. M. Linnanto and J. E. I. Korppi-Tommola, Photosynth. Res., 2008, 96, 227 CrossRef CAS PubMed.
  37. B.-J. van Rossum, D. B. Steensgaard, F. M. Mulder, G. J. Boender, K. Schaffner, A. R. Holzwarth and H. J. M. de Groot, Biochemistry, 2001, 40, 1587 CrossRef CAS PubMed.
  38. V. Huber, M. Katterle, M. Lysetska and F. Würthner, Angew. Chem., Int. Ed., 2005, 44, 3147 CrossRef CAS PubMed.
  39. M. Gouterman, J. Mol. Spectrosc., 1961, 6, 138 CrossRef CAS.
  40. M. Gouterman and G. H. Wagniere, J. Mol. Spectrosc., 1963, 11, 108 CrossRef CAS.
  41. The emission lifetime of the chiral J-aggregates of ZnChls could not be determined because it was shorter than the instrument response function.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2tc04841h

This journal is © The Royal Society of Chemistry 2023