Kyohei
Nakano
a,
Iat Wai
Leong
ab,
Daisuke
Hashizume
a,
Kirill
Bulgarevich
a,
Kazuo
Takimiya
acd,
Yusuke
Nishiyama
e,
Toshio
Yamazaki
f and
Keisuke
Tajima
*a
aRIKEN Center for Emergent Matter Science (CEMS), 2-1 Hirosawa, Wako 351-0198, Japan. E-mail: keisuke.tajima@riken.jp
bSANKEN, Osaka University, Mihogaoka 8-1, Ibaraki, Osaka 567-0047, Japan
cDepartment of Chemistry, Graduate School of Science, Tohoku University, 6-3 Aoba, Aramaki, Aoba-ku, Sendai, Miyagi 980-8578, Japan
dTohoku University Advanced Institute for Materials Research (AIMR), 2-1-1 Katahira, Aoba-ku, Sendai, Miyagi 980-8577, Japan
eJEOL Ltd., Musashino, Akishima, Tokyo 196-8558, Japan
fRIKEN Center for Biosystems Dynamics Research, 1-7-22 Suehiro-cho, Tsurumi-ku, Yokohama, Kanagawa 230-0045, Japan
First published on 16th October 2023
To investigate potential applications of the 3,3′-dihydroxy-2,2′-biindan-1,1′-dione (BIT) structure as an organic semiconductor with intramolecular hydrogen bonds, a new synthetic route under mild conditions is developed based on the addition reaction of 1,3-dione to ninhydrin and the subsequent hydrogenation of the hydroxyl group. This route affords several new BIT derivatives, including asymmetrically substituted structures that are difficult to access by conventional high-temperature synthesis. The BIT derivatives exhibit rapid tautomerization by intramolecular double proton transfer in solution. The tautomerizations are also observed in the solid state by variable temperature measurements of X-ray diffractometry and magic angle spinning 13C solid-state NMR. Possible interplay between the double proton transfer and the charge transport is suggested by quantum chemical calculations. The monoalkylated BIT derivative with a lamellar packing structure suitable for lateral charge transport in films shows a hole mobility of up to 0.012 cm2 V−1 s−1 with a weak temperature dependence in an organic field effect transistor.
Indigo, an organic dye with a long history, has a π-conjugated structure with intramolecular hydrogen bonds formed between two sets of N–H and CO groups (Fig. 1a). The compound also forms an intermolecular hydrogen bonding network in the solid state, which renders indigo virtually insoluble. Recently, indigo and its derivatives have been used as OSCs with decent ambipolar mobility (0.56 cm2 V−1 s−1 for hole and 0.95 cm2 V−1 s−1 for electron) and high stability toward various environments.5–8 3,3′-Dihydroxy-2,2′-biindan-1,1′-dione (BIT; 1 in Fig. 1b) has been known as a structural analogue of indigo for more than a century,9,10 but it has never been used as an OSC. BIT also has two strong intramolecular hydrogen bonds between the enolic O–H and CO, resulting in the planar structure of the molecules. 1 has an intense purple color, suggesting the delocalization of electrons through π-conjugation. In addition, unlike indigo, 1 exhibits tautomerization by the simultaneous transfer of two protons from O–H to CO to reproduce the mirror inverted hydrogen-bonded enol structures (Fig. 1b). These mobile protons within the tautomeric molecules are a unique feature of BIT compared to indigo or other hydrogen-bonded OSCs. Tautomerization of polar molecules by proton transfer in organic solids can switch the direction of the polarity and has been extensively studied in relation to the dielectric and ferroelectric properties of insulating materials.11–16 Even without the switching of polarity, tautomerization can also affect the molecular orbitals and the intermolecular electronic coupling in the aggregates that determine the charge transport properties in the solid state. If rapid tautomerization can be realized in OSCs, the interplay between the intramolecular proton transfer and the electronic transport could lead to a new class of OSCs with unique dynamic physical properties.17
Fig. 1 (a) Structure of indigo and (b) identical tautomeric structures of 3,3′-dihydroxy-2,2′-biindan-1,1′-dione (1). |
To explore the possibility of BIT as an electronic functional material, we need various derivatives of the structure. However, most of the BIT derivatives reported so far involve reactions of the keto–enol structures, which loses the intramolecular hydrogen bonds and the planarity of the molecules. For example, the Wudl group reported the synthesis of zwitterionic BIT derivatives with cyanine and dione enolate structures and their application as OSCs in organic field effect transistors (OFETs) with a hole mobility of 2.1 × 10−4 cm2 V−1 s−1.18 To our knowledge, the use of BIT-related structures for electronic transport has been limited to this report, mainly because the available synthetic route for the BIT structure requires harsh conditions at high temperatures, and the reaction has been limited to non-substituted or symmetric and simple substitutions on the benzene rings.19,20
To elucidate the potential of the BIT structure in OSCs, in this study, we present a new synthetic route for BIT consisting of the addition reaction of 1,3-dione to ninhydrin followed by hydrogenation of the hydroxyl group at room temperature. This route allows us to access various BIT derivatives with the intact hydrogen bonding moieties and asymmetric substitutions. The tautomerizations involving the double proton transfer are studied both in solution and in the solid state. The possible interplay between the proton transfer and the charge transport properties is also studied by quantum chemical calculations. Some of the alkylated BIT derivatives with suitable packing structures for charge transport in films are tested in OFETs.
Fig. 3 Structure of BIT derivatives (2–12) with symmetric and asymmetric substitutions synthesized by the new route shown in Fig. 2b. The numbers in parentheses are the isolated yields starting from the corresponding indan-1,3-diones and the ninhydrins. |
To overcome these problems, we developed the synthetic route shown in Fig. 2b. First, 2-hydroxy-2,2′-biindan-1,1′,3,3′-tetraone was synthesized by the addition reaction of indan-1,3-dione to ninhydrin in acetic acid at room temperature in moderate yields (40–60%).22 In the second step, the hydrogenation of –OH with poly(methylhydrosiloxane) (PMHS) catalyzed by methanesulfonic acid or trifluoromethanesulfonic acid afforded 1. Under the reductive conditions tested (i.e., hydrosilanes with Et2O·BF3 or B(C6F5)3, Pd/C and H2, and formic acid23), PMHS/protonic acids gave the best product yields (30–70%).24 Et3SiH also worked as a hydrogen source in the combination of the protonic acids, but gave a lower yield than PMHS. Although the total yield of 1 obtained with the new route (17% after the two steps) was not substantially higher than that obtained with the Perkin route, both the reaction steps in Fig. 2b were performed at room temperature, which prevented the formation of the byproduct. The milder conditions also allowed the synthesis of BIT derivatives with various functionalities and to achieve the asymmetric substitution of BIT starting from either or both indan-1,3-diones or ninhydrins functionalized on the benzene rings.
The applicability of the new synthetic route was tested with several derivatives of indan-1,3-dione and ninhydrin. The reaction between 5-alkyl-1,3-indandione and ninhydrin afforded BIT derivatives with asymmetric alkyl substitutions (3–6 in Fig. 3). 2 was synthesized by the reaction between 5-methyl-1,3-indandione and 4-methylninhydrin to give the product as a mixture of two structural isomers (5,5′-dimethyl and 5,4′-dimethyl, see NMR Spectra in solution section below). Reactions of tetrafluoroindan-1,3-dione with ninhydrin or 5-alkylated ninhydrin afforded the corresponding asymmetrically fluorinated (7) or fluorinated and alkylated (8) derivatives. Perfluorinated BIT (9) was also synthesized from tetrafluoroindan-1,3-dione and tetrafluoroninhydrin by the same reactions but in a lower yield. This is because the tetrafluoroninhydrin could not be isolated, probably due to its low stability, and the second steps of the reactions in Fig. 2b were carried out without isolating the product from the first step. Benz[f]indan-1,3-dione was also used in the reactions with ninhydrin to give the asymmetric compound (10). Benz[f]indan-1,3-dione and benzo[f]ninhydrin were used to synthesize the symmetric 3,3′-dihydroxy-2,2′-bibenz[f]indan-1,1′-dione (11) and the asymmetric 3,3′-dihydroxy-2,2′-benz[f]indan-5-indan-1,1′-dione (12). The synthetic details and characterizations are presented in the ESI (Fig. S1–S6).†
Considering the large variety of indan-1,3-dione analogues available and the rich chemistry of ninhydrin,25–27 these results demonstrate that the new synthetic route in Fig. 2b can be used to access various substituted BIT derivatives that cannot be easily synthesized by the conventional route in Fig. 2a.
Fig. 4 1H NMR spectra of 1–12 in the region of the enol –OH protons. Solvents are CDCl3 (25 °C) for 1–8 and 12 and 1,1,2,2-tetrachloroethane-d2 (120 °C) for 9–11. |
The 13C NMR spectra of the solutions also indicated the rapid conversion between the tautomers by proton transfer (Fig. 5). The spectra for 1 and 9 showed a single peak at 187.58 and 182.24 ppm, respectively, corresponding to CO and C–OH, indicating that the four carbons connected to O were equivalent. 3–6, 8, and 12 showed four peaks corresponding to CO and C–OH, instead of the eight peaks expected for the mixture of the tautomers with nonequivalent CO and C–OH carbons. 7 and 10 showed two peaks for CO and C–OH, instead of four peaks expected for nonequivalent carbons with fixed –OH positions.
Fig. 5 13C NMR spectra of 1–10 and 12 in the region of the CO and C–OH carbons. Solvents are CDCl3 (25 °C) for 1–8 and 12 and 1,1,2,2-tetrachloroethane-d2 (120 °C) for 9–10. |
The 1H NMR spectrum of 2 shows two peaks for the 5,4′-dimethyl isomer (14.69 ppm and 14.53 ppm) and one for the 5,5′-dimethyl isomer (14.59 ppm), and the ratio of the isomers was approximately 1:1 based on the integrations. Our attempts to separate the isomers of 2 failed, suggesting isomerization of the molecules in solution. The formation of small amounts of the diketo tautomers is likely to cause a slow conversion between the isomers, while rotation of the central C–C bond was unlikely due to the intramolecular hydrogen bonds. In support of this, when variable temperature 1H NMR was performed for 4, the two –OH peaks showed no sign of coalescence up to 170 °C (Fig. S11†), suggesting that the environment for the two protons remained nonequivalent at high temperatures.
To investigate the properties of the charged species that would be formed in OFETs, the molecular structures of the radical cation (1˙+) and the radical anion (1˙−) are optimized in the ground and the transition states (Fig. 7). The ground state of 1˙+ have the structure with C2h symmetry similar to 1 with a slightly longer O–H bond length (1.02 Å) than 1 (1.01 Å) (Fig. 7a). The transition states of 1˙+ is also similar to that of 1 with C2v symmetry with off-centered protons and a shorter O–H distance of 1.08 Å (Fig. 7b). The barrier energy of double proton transfer for 1˙+ is 0.12 eV, which is much lower than that for 1. This implies that the protons are more mobile in the positively charged molecules than in the neutral ones. In contrast, 1˙− has significantly different structures from 1 in the ground state with C2v symmetry and the two protons positioned on one of the sides with an O–H bond length of 1.01 Å (Fig. 7c). This resulted in the formation of a large dipole moment of 3.0 Debye in the direction of the long axis of the molecules. The transition states of 1˙− have C2h symmetry with a shorter O–H distance of 1.03 Å (Fig. 7d). In contrast to 1˙+, 1˙− has a higher barrier energy of the transition states (0.34 eV) than 1. This implies that the protons are less mobile in the negatively charged molecules.
The above calculations are also performed for 2, 3, 7, 8, 9, 10, 11, and a model of 12. The tautomers are defined as in Fig. S12–S14† and the calculated results are summarized in Tables S1–S4.† The electronic energy and the HOMO and LUMO energies are almost the same between the tautomers even for the BITs with asymmetric structures (Table S1†). The barrier energy of the double proton transfer for the neutral molecules, the radical cations, and the radical anions are in the range of 0.17–0.23 eV, 0.06–0.13 eV, and 0.33–0.46 eV, respectively (Tables S2–S4†). There was a general tendency for the barrier energy of radical anions, neutral molecules, and radical cations to be in descending order for each molecule.
Fig. 8 Absorption spectra of (a) 1 in CHCl3 solution and (b) 4 in a 100 nm-thick film prepared by vacuum deposition. |
Fluorescence spectroscopy of 1 in o-dichlorobenzene solution showed no observable fluorescence up to 850 nm, indicating that the BIT structure is intrinsically non-fluorescent. We speculate that the reason for the absence of the fluorescence could be related to the excited state proton transfer process in BIT, although the detailed photophysical behaviors require further investigation. In contrast, 1 synthesized with the Perkin reaction showed strong fluorescence in o-dichlorobenzene solution at 575 nm with vibronic features (Fig. S17†). The excitation spectra also had vibronic features that did not match the broad absorption spectrum around 500 nm. This mismatch between the absorption and the excitation spectra indicated that 1 synthesized with the Perkin reaction contained a fluorescent impurity, which was probably a structural isomer formed during the high-temperature synthesis. We failed to remove this fluorescence impurity by repeated recrystallization or vacuum sublimation. Avoiding this impurity is another advantage of our new synthetic route.
The ionization energy of 6 was 5.56 eV, determined by photoemission yield spectroscopy in air on a thin film prepared by vacuum deposition (Fig. S18†), which was consistent with the energy level of HOMO estimated by the DFT calculations (−5.63 eV). The ionization energy of BIT is deeper than that of a typical p-type OSC, such as [1]benzothieno[3,2-b]benzothiophene (5.45 eV)30 or pentacene (4.94 eV),31 but hole injection from the electrode is possible for BIT derivatives.
1 | 2 | 7 | 8 | 9 | 11 | 12 | |
---|---|---|---|---|---|---|---|
a There are two other O–H bonds (O1–H and O3–H) due to the packing disorder and the average values are presented. | |||||||
C1–O1 | 1.2549(9) | 1.2659(18) | 1.2326(5) | 1.228(4) | 1.231(4) | 1.24750(19) | 1.267(4) |
C3–O3 | 1.3019(9) | 1.2948(18) | 1.3091(5) | 1.306(4) | 1.311(4) | 1.30740(19) | 1.296(4) |
C1–C2 | 1.4482(10) | 1.441(2) | 1.4645(5) | 1.465(4) | 1.464(4) | 1.4563(2) | 1.441(5) |
C2–C3 | 1.4033(10) | 1.415(2) | 1.3896(5) | 1.385(4) | 1.389(4) | 1.3964(2) | 1.418(5) |
C2–C2′ | 1.4606(9) | 1.454(3) | 1.4549(5) | 1.458(4) | 1.455(6) | 1.4566(3) | 1.470(5) |
C1′–O1′ | 1.2567(9) | 1.2659(18) | 1.2357(4) | 1.238(4) | 1.231(4) | 1.24750(19) | 1.269(4) |
C3′–O3′ | 1.3006(9) | 1.2948(18) | 1.3056(5) | 1.303(4) | 1.311(4) | 1.30740(19) | 1.291(4) |
C1′–C2′ | 1.4482(10) | 1.441(2) | 1.4620(5) | 1.458(4) | 1.464(4) | 1.4563(2) | 1.447(5) |
C2′–C3′ | 1.4003(10) | 1.415(2) | 1.3921(5) | 1.388(4) | 1.389(4) | 1.3964(2) | 1.414(4) |
O1–O3′ | 2.5418(9) | 2.5343(18) | 2.5642(4) | 2.568(3) | 2.548(3) | 2.5162(2) | 2.534(3) |
O3–O1′ | 2.5274(8) | 2.5343(18) | 2.5470(4) | 2.549(3) | 2.548(3) | 2.5162(2) | 2.532(3) |
O3–H | 1.09(3) | 0.91a | 0.980(19) | 0.95(5) | 1.00(5) | 1.030(7) | 0.96(5) |
O3′–H | 1.07(2) | 0.91a | 0.985(18) | 0.97(4) | 1.00(5) | 1.030(7) | 1.03(6) |
In the molecular structure of 1 in the crystal (Fig. 9a and Table 1), at the first glance, the tautomeric structures and the proton positions of –OH are fixed with static conjugated double bonds; the bond lengths of C3–O3 and C1–C2 were longer than those of C1–O1 and C2–C3, respectively. However, the observed length differences between CO and C–OH (0.045 Å) and between CC and C–C (0.046 Å) are smaller than the calculated bond length differences in 1 by DFT (0.075 Å and 0.077 Å, respectively, see Table S7†). The observed smaller length differences suggest that there may be either dynamic or static disorder in term of the tautomeric structures in the crystal of 1.
As indicated by 1H NMR, 2 contained the cis and trans isomers in a 1:1 ratio in solution. In the single crystal, the methyl groups had an occupancy of about 0.5 at the four possible positions (not shown in Fig. 9b for clarity). This suggests that 2 shows structural disordering due to both the orientation of the molecules and the co-crystallization of the two isomers. In this case, this static packing disorder could contribute to the observed small differences in bond length between CO and C–OH (0.029 Å) and those of CC and C–C (0.026 Å).
In the crystal structure of 7 (Fig. 9c), the –OH protons occupied a single position, and the bond length differences between CO and C–OH (0.074 Å) and between CC and C–C (0.073 Å) are close to the calculated values (0.075 Å and 0.076 Å, respectively). In the molecular packing of 7, the fluorinated and non-fluorinated benzene rings were closely stacked together, suggesting the presence of donor–acceptor interactions between them. The crystal structure of 8 was also similar to that of 7 (Fig. 9d), with a dominant contribution of one of the tautomers and similar packing motifs, but there was disorder in the molecular arrangement in the direction of the methyl substituents, with the two directions present in approximately a 1:1 ratio.
The crystal structure of 9 was a layered structure with in-plane molecular packing in a herringbone pattern (Fig. 9e), and the contribution from one of the tautomers was dominant as for 7 and 8 at 90 K. The distance between π-planes was the shortest among the compounds (3.13 Å, Table S6†). In the crystal structure of 11 (Fig. 9f), the intramolecular O⋯O distance was the shortest among the compounds (2.52 Å, Table 1), and the bond length differences between CO and C–OH (0.060 Å) and between CC and C–C (0.060 Å) are slightly smaller than the calculated values (0.069 Å and 0.068 Å, respectively). The molecular packing was a pitched π-stacking structure with a large slip in the long axis direction. This structure was similar to the crystal of indigo, although unlike indigo, there are no intermolecular hydrogen bonds in 11.33
The crystal structure of 12 was a layered structure with antiparallel alignment of the molecules in in-plane molecular packing (Fig. 9g). The bond length differences between CO and C–OH (0.026 Å) and between CC and C–C (0.028 Å) are smaller than the calculated values (0.071 Å and 0.070 Å, respectively). The distance between π-planes was 3.41 Å, but the overlap of the π-conjugated core in the crystal is small due to the large displacement.
We are interested in whether the double proton transfer also occurs in the solid state. To investigate the dynamic disorder due to tautomerization, the single-crystal X-ray analysis was performed at different temperatures and the changes in bond length are analyzed. The results for 1 and 7 are summarized in Tables S8 and S9,† respectively. As the measurement temperature is increased from 90 K, the bond length differences between CO and C–OH and between CC and C–C become smaller (0.015 Å and 0.013 Å, respectively, for 1 at 373 K) to give a more symmetrical structure around the central part of the molecules. This result suggests that the dynamic tautomerization process occurred in the solid state to give the averaged structure of two tautomers in the X-ray analysis.
To visualize the position of the hydrogen atom more directly, the electron density distribution was examined by taking the difference Fourier maps on the plane average calculated using the structure models without the H atoms on the –OH groups and the measurements, where the protons appear as positive electron density. The results for 7 are shown in Fig. 10. The electron density difference corresponding to the proton in –OH is observed at the two positions between the O⋯O with the electron density ratio of about 77:23 at 90 K. When the measurement temperature was increased to 298 K and 423 K, the density ratio of the proton changed to 65:35 and 58:42, respectively. These results indicate that the tautomerization occurred in the solid state by intramolecular proton transfer. The similar tendency was observed for 1 in variable temperature measurements (Fig. S19†). The molecular C2h symmetry of 1 leads to the symmetric double-well potential for the proton as shown by the DFT calculations. However, we speculate that the molecular packing in the single crystals makes the double well potential asymmetric. This leads to the distorted population ratio of the tautomers at thermal equilibrium depending on the temperatures according to the Boltzmann distribution. The energy difference between the tautomers in the crystal of 7 was calculated to be about 1.4 kJ mol−1 from the population ratios at 298 K and 423 K, assuming that the system had reached thermal equilibrium.
To further elucidate the dynamic tautomerization process in the solid state, we have performed magic angle spinning (MAS) 13C solid-state NMR of 1 (the full spectra are shown in Fig. S20 and S21†). The spectrum shows two peaks at 190.15 ppm and 186.28 ppm at 298 K, which can be assigned to CO and C–OH, respectively (Fig. 11). To estimate the chemical shift difference in the solid state with the fixed tautomeric structures (at 0 K), we have performed calculations based on the Gauge Including Projector Augmented Waves (GIPAW) method, using the crystal structure of 1 as the starting structure.34–36 The details of the calculations are given in the ESI.† The results show that the difference in the chemical shifts between CO and C–OH in the crystal is expected to be 15.64 ppm. The observed chemical shift difference (3.87 ppm) is much smaller than the calculated value, indicating that the two 13C environments are rapidly exchanged by tautomerization. The chemical shift difference reflects the biased population ratio of the tautomers at thermal equilibrium. Using the calculated chemical shift difference as the reference for the fixed tautomeric structures with an estimated error of ±2 ppm,34 the population ratio of the tautomers can be calculated from the observed values to be in the range of 0.64:0.36 and 0.61:0.39 at 300 K and the energy difference between two tautomers can be estimated to be 1.1–1.4 kJ mol−1. These results were in agreement with the population ratios estimated from the single-crystal X-ray analysis. When the measurement temperature was increased to 423 K, the difference in the chemical shift became smaller (Fig. 11), indicating the change in the population ratio of the tautomers according to the Boltzmann distribution. These results further support the picture that the fast double proton transfer in 1 occurs above the room temperature within the asymmetric double well potential induced by the crystal structure.
Fig. 11 Region of CO and C–OH peaks in the magic angle spinning (MAS) 13C solid-state NMR spectra of 1 measured at 298 K (black) and 423 K (red). |
Fig. 12 Two tautomeric situations A and B used to calculate the transfer integrals in the pair of 1. The colored lines indicate the difference in proton positions. See the ESI† for details on the procedures used to generate the structures. |
1 has a relatively large reorganization energy of 0.41 eV for the hole. This could be attributed to the structure change of BIT with the slight shift of the proton positions in 1˙+ (Fig. 7a). On the other hand, the reorganization energy of 1 for the electron is very large, about 1.07 eV. This could be attributed to the large structural change in the ground state of 1˙− in which the two protons are shifted to one side of the molecules as shown in Fig. 7c. Therefore, it is expected that the electron transport in BITs could be strongly prohibited.
The calculations of the transfer integrals for the pairs and the reorganization energy are also performed for 2, 7, 8, 9, and 11, and the results are summarized in Fig. S22 and Tables S10–11.† The difference in the transfer integrals between the two tautomeric situations was observed for most of the molecules (7–38%). The large reorganization energy for the charge transfer (0.36–0.48 eV for the hole and 0.96–1.24 eV for the electron) is a common feature in the BIT derivatives.
The surface morphology of the film of 6 observed by atomic force microscopy (AFM) is shown in Fig. 14c. A step and terrace structure with a height of about 7.8 nm was observed, which was almost twice the spacing of the 001 reflection in the out-of-plane direction (3.8 nm). The lateral sizes of the steps and the terraces were about 1 μm, which was small compared to those of state-of-the-art small-molecule semiconductors with high mobility (>10 μm).37 AFM images for the vacuum deposited films of 1, 2, 4, 5, 9, 11, and 12 are shown in Fig. S25.† The films of 1 and 2 consist of large needle-shaped crystals with a length of several μm, and the poor connectivity of the domains may not be suitable for OFET applications. The films of 4 and 5 showed step-terrace features on the surface as in 6, suggesting the formation of uniform layered structures. The films of 9, 11 and 12 show the presence of crystalline grain structures without clear step-terrace features, which may be related to the lack of charge transport properties in OFETs. Further optimization of the film structures may be necessary to improve the charge mobility of BITs in OFETs.
The temperature dependence of the hole mobility was studied for 6 in OFETs. In the range of 25 °C and 90 °C, the mobility showed a weak dependence on temperature; the higher the temperature, the lower the hole mobility (Fig. S26†). This is in contrast to typical thermally activated charge transport in disordered OSCs, where the charge hopping is activated by thermal energy. Whether this behavior is related to the interplay between intramolecular double proton transfer and charge conduction is unknown at this time and will be investigated in the future. The device showed severe degradation above 100 °C due to evaporation of the films in vacuum.
Footnote |
† Electronic supplementary information (ESI) available: Additional experimental details on materials, instruments, quantum chemical calculations, device fabrication and characterizations, and synthetic procedures, crystallographic data as CIF files, 1H NMR, UV-vis and FL spectra, and photoemission yield spectra. CCDC 2216887–2216892 and 2278636. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3sc04125e |
This journal is © The Royal Society of Chemistry 2023 |