Introduction
Organic light-emitting diodes (OLEDs) are widely used in mobile devices such as smartphones, flat panel displays, and wearables. They are lightweight, ultrathin, wide-angle, and flexible and can provide an excellent viewing experience. As the color gamut standard for next-generation displays, the ITU-R BT.2020 standard requires the full width at half-maximum (FWHM) of the monochromatic red, green, and blue (RGB) spectra to be less than 30 nm.[1–4] Conventional organic emitters, such as perylene and coronene, have very rigid polycyclic aromatic hydrocarbons and commonly exhibit relatively small structural relaxations with small Stokes shifts. However, the “breath” of their excited state is produced mainly by the vigorous stretching vibration of the molecular bond, producing multiple fine vibration emissions with large FWHM values.[5–7] In 2016, Hatakeyama et al. developed a N/B-doped rigid heterocycle,[8] in which the frontier orbitals were effectively separated. This so-called multiple resonance (MR) effect could attain short-range charge transfer transition and result in a small FWHM value of the EL spectrum.[2,9–13] Nevertheless, MR emitters have two key issues including the evident aggregation-caused photoluminescence (PL) quenching and spectral broadening.[14–17] For example, Monkman et al. reported that even at extremely low doping concentrations, MR emitters could exhibit a shoulder peak at a longer wavelength, resulting in large FWHM values and lower-than-expected device performance.[17,18]
The FWHM value is dependent on the structural relaxation and vibration coupling of the excited states. Molecular vibration is closely associated with the optical phonons, as described by Marcus theory. This connection implies that the interaction between the excitons and optical phonons can manifest as exciton-vibration coupling. To quantify this interaction, the FWHM of the PL spectra can be fitted to the following equation (Equation (1)):[19]
Here, Γ(T) represents the FWHM at temperature T, and S and ħω denote the Huang–Rhys factor and the effective vibrational energy coupled in the emission spectra, respectively. The FWHM of PL spectra generally increases along with the molecular vibration due to the strong interaction between the excitons and optical phonons. In the context of aggregation-induced PL quenching, the photoluminescence quantum yields (PLQYs) of organic emitters are determined by the nonradiative decay rate constant (knr). According to the band-gap law, knr can be simply expressed as follows:[20,21]
Pressure, as an independent thermodynamic parameter, allows intermolecular interactions to be precisely manipulated. Moreover, it drastically alters the electronic structure and bond vibrations of materials and, subsequently, their photophysical properties.[22–25] Recently, the application of pressure using the diamond anvil cell (DAC) technique, in combination with spectroscopic detection, has provided valuable insights into the structure‒property relationship.[26–28] Herein, we follow the structural changes at the molecular level and correlate the subtle alterations in molecular geometry and intermolecular distances with the photophysical properties of the MR emitters using a combination of high-pressure in situ optical spectroscopy and theoretical calculations. This combination enables the elucidation of the effect of exciton‒vibration couplings on the FWHM and emission efficiency of the MR emitters in the solid state; this aspect is currently poorly understood. The narrowband MR emitter (BNC, Figure 1a) is prepared by incorporating spherical o-carborane (C2B10H12) as steric hindrance group. The incorporation of icosahedral boron clusters slightly elevates the high-frequency vibration of the MR framework due to the absence of C═C bonds while simultaneously weakening the molecular interactions. Accordingly, the boron cluster BNC shows a PLQY of 92.9% and small FWHMs of 29 nm at low dopant concentrations. In the crystalline state, the MR-emitting molecules, which assemble into “isolated dimers, ”clearly quench the PL, with a PLQY of 15.0%. Under low pressure, the PL spectra of BNC crystals exhibit a noticeable broadening, which is primarily attributed to the significant weakening of the PL peak at 536 nm. Interestingly, the low-frequency and partially high-frequency vibrations are consistently suppressed from 0.48 to 4.43 GPa, resulting in a reduction in the exciton‒vibration couplings. Consequently, BNC crystals exhibited a noticeable increase in intensity and a slight narrowing of the PL spectra, which are rarely observed in previous reports.[29–31] At pressures above 4.50 GPa, the molecular distance between the “isolated dimers” decreases, leading to the appearance of excimer fluorescence. The excitonic interactions between closely stacked “isolated dimers” cause fluorescence weakening and PL spectral broadening. These data not only demonstrate the exciton‒vibration coupling behavior of this particular MR emitter at high pressure but also establish that high-pressure optical techniques provide rich structure‒fluorescence relationships.
[IMAGE OMITTED. SEE PDF]
Results and Discussion
Photophysical Properties and Structure of BNC under Ambient Conditions
The photophysical properties of BNC are shown in Figure 1, along with its ultraviolet–visible (UV–vis) absorption and PL spectra. As depicted in Figure 1a, the UV‒Vis spectrum of toluene (1 × 10−5 m) exhibited the characteristic MR absorption profile: the intense absorption band at approximately 284 nm corresponded to the local π–π* transitions, and the sharp absorption peak at 471 nm was attributed to the charge transfer transition.[32] The BNC solution emitted bright green fluorescence with a PLQY of up to 83.5% (Figure 1b). Its PL spectra demonstrated a clear peak at 492 nm with a very small FWHM of 23 nm. A small FWHM of 22 nm was also observed for the reference luminophore BNH (Figure S4, Supporting Information), which did not contain any spherical o-carborane units. These results showed that the MR characteristics of the B-N polycyclic framework were not interrupted by icosahedral o-carborane. Moreover, the density functional theory (DFT) calculations were performed at the B3LYP/6-31G(d,p) level. As depicted in Figure S5 (Supporting Information), BNC and BNH had nearly identical nonbonding molecular orbital distributions, which were consistent with the negligible change in the FWHM of BNC relative to BNH. Interestingly, the PL spectrum showed a mirror image correlation with the absorption, and the observed Stokes shift was only 21 nm. The similarity between the spectra of the crystalline powder, as shown in Figure 1a,b confirmed that the small Stokes shift was caused by the slight structural deformation rather than solvent relaxation.[33] Remarkably, the PL spectrum of BNC in the crystalline state had a shoulder peak at approximately 587 nm compared with that in toluene. The transient PL decay curves revealed that the lifetime of the shoulder peak (8.8 ns, Figure 1c) was longer than that of the main peak (1.4 ns). This difference, combined with the molecular packing (Vide infra, Figure 2a), strongly demonstrated that the fluorescence peak at 587 nm was due to the excimer state. In summary, the narrow FWHM and small Stokes shift of BNC were attributed to the unique distribution of the frontier molecular orbitals on the BNC core, which was a result of the nonbonding characteristics of the highly rigid MR framework. This distinctive feature essentially reduced the vibronic coupling between the S0 and S1 states, as well as the vibrational relaxation at the S1 state.
[IMAGE OMITTED. SEE PDF]
To better understand the origin of the shoulder peak (Vide infra), the BNC crystal was meticulously grown in chloroform through solvent diffusion. The corresponding structures were determined using single-crystal X-ray diffraction experiments and belonged to a monoclinic crystal with lattice parameters of a = 16.6148(11) Å, b = 18.6620(12) Å, c = 19.4867(13) Å, α = 99.843(2)°, β = 113.550 (2)°, and γ = 115.882(2)°. In the crystalline state, the BNC monomers had relatively planar structures with a molecular planarity parameter (MPP) of 0.780 Å. The molecules of the BNC crystals adopted two distinct kinds of antiparallel arrangements (Figure 2a): the tail-to-tail and head-to-tail patterns. For the former, the interplanar π−π distance (Dtt) between the two MR units was 3.47 Å, and the overlapping area was greater than 50% (Figure 2b,c). Moreover, the energy of the interplanar interaction was calculated to be −251 kJ mol−1. Therefore, the occurrence of the strong π‒π interaction likely led to the formation of antiparallel dimers. In contrast, one molecule of BNC slid away from its neighbor along the long axis of the molecule, as shown in Figure 2d. The distance between two boron atoms increased from 4.76 Å in the tail-to-tail dimer (Figure S6, Supporting Information) to 10.90 Å, resulting in head-to-tail packing. Figure 2d shows that the dihedral angle (θpmr) between the phenyl ring and its adjacent MR framework was 31.1°. Moreover, the distance Dth between the phenyl centroid of BNC and the plane of the adjacent molecule reached 4.79 Å (Dth, Figure 2e). These observations indicated the absence of the interplanar π‒π interactions in the head‒to-tail packing. Therefore, this antiparallel tail-to-tail arrangement could be regarded as the isolated and discrete π‒π dimer stacking in the BNC crystals. These results, together with the longer PL lifetime (8.8 ns, Vide supra), showed that the shoulder peak of the BNC crystals at 587 nm was probably due to excimer luminescence of the “isolated dimer” structure. Generally, the “isolated dimer” packing ensures the purity and singularity of the excimer state through exciton localization, thereby preventing the formation of a “dark” state.[34,35] Thus, unique dimeric stacking, which promoted discrete excimer formation, was extremely beneficial for achieving high-efficiency fluorescence. However, in our case, the PLQY of the BNC crystals was 15.0%, which was significantly lower than that of the PMMA film (1‰ wt/wt, PLQY = 92.9%, Figure S7, Supporting Information). This aggregation-induced PL quenching could be attributed to the high nonradiative rate constant (knr = 1/τ − kr) of 0.6 ns−1. Then, we performed temperature-dependent PL spectral analysis on the crystalline BNC powder. As the temperature decreased, the PL peak exhibited a blueshift and increased in intensity (Figure S8a, Supporting Information); similar results were previously observed in nonfullerene acceptors.[36] Concurrently, the integrated PL intensity monotonically decreased with increasing temperature, as illustrated in Figure S8b (Supporting Information); this enabled the derivation of the corresponding exciton binding energy (Eb) as 34.9 meV using the Arrhenius equation,[37] I(T) = I0/[1+Aexp(-Eb/kbT)], where I0 represents the integrated intensity at 0 K, kb is the Boltzmann constant and Eb denotes the exciton binding energy. Notably, when BNC was doped in PMMA at a weight concentration of 1 wt‰, the resulting film exhibited a large Eb of 72.4 meV (Figure S9, Supporting Information), nearly double that of the BNC crystals. Clearly, in the doped film, the high Eb ensured the generation of excitons at room temperature and their high-rate recombination, and hence, a high PLQY was attained. From the doped film phase to the crystals, the Eb demonstrated a significant decrease due to the strong polarization effects[38] and was slightly greater than the thermal disturbance energy (≈26.0 meV) at room temperature. Therefore, the spontaneous dissociation of excitons at room temperature was effortless, ultimately causing aggregation-induced PL quenching.
Photophysical Properties of MR BNC at High Pressure
The photophysical properties of BNC changed in response to the application of high pressure, as the molecular packing in the crystals became more compact. As shown in Figure S10a (Supporting Information), piezochromism was visible to the naked eye, with the color of the crystals changing from yellow to deep red as the absorption spectrum shifted toward the red region. The UV‒visible absorption red edge progressively shifted from 518 nm at ambient pressure to 635 nm at 10.20 GPa, resulting in a bathochromic shift of 3557 cm−1. Accordingly, the optical energy gap shifted from approximately 2.43 to 2.02 eV (Figure S10b, Supporting Information), indicating that the energy gap of the BNC decreased during compression. When high pressure was released, the UV spectrum reverted back to its original state (Figure S11a, Supporting Information). Interestingly, a good linear relationship was observed between the optical energy gap and external pressure (Figure 3f). Below 0.48 GPa, the PL intensity of the BNC crystals was slightly weakened (Figure 3a,e), in contrast to the sharp broadening of the PL spectrum. When the applied pressure was gradually increased from 0.48 to 3.43 GPa, the PL intensity of BNC significantly and rapidly increased (Figure 3b). As illustrated in Figure 3e, the PL intensity at 3.43 GPa exhibited a remarkable 50% increase relative to that at 0.48 GPa. Moreover, the FWHM of the PL spectra was slightly narrowed, as shown in Figure 3d. Above 4.50 GPa, BNC crystals showed gradually broadened PL spectra and decreased intensity, with a slight redshifted fluorescence (Figure 3c,e). Clearly, under high pressure, the redshift of the absorption and PL spectra clearly indicated a decrease in the energy of the low-lying excited state. These results indicated that high pressure induced a more compact arrangement of molecules, resulting in increased intermolecular interactions.[25,27,28] Upon pressure release, the fluorescence peak returned to its original peak at 536 nm; however, the intensity of the shoulder peak at 587 nm significantly increased (Figure S11b, Supporting Information). As shown in Figure S12 (Supporting Information), the decompressed BNC crystals displayed a single exponential PL decay curve and had a longer lifetime than the original state. Thus, the strength of the interplanar π‒π interactions within the “isolated dimers” did not completely return to their initial state.
[IMAGE OMITTED. SEE PDF]
Structure of MR BNC at High Pressure
Theoretical calculations of the BNC crystals at various pressures were performed to clarify the abnormal piezochromic fluorescence. Upon compression of the BNC from ambient pressure to 9.00 GPa, the unit cell volume contracted by 722 Å3 (−28.0%) (Figure S13a, Supporting Information). The compression of the unit cell was anisotropic, and the a-, b-, and c-axes were shortened by 7.33 Å (−26.0%), 1.51 Å (−6.8%), and 0.926 Å (−4.4%), respectively. The most compressible axis (the a-axis) was aligned with the decrease in the interplanar distance Dtt between two MR units (Vide infra, Figure 4b), whereas the less compressible b- and c-axes were coincident with slight slip along the long axis of the molecule. In Figure 4a,d, the molecular coplanarity greatly improved upon compression, as evidenced by a 62.7% decrease in the dihedral angle, θpmr, from 31.1° to 11.6°. Furthermore, the distance between neighboring boron atoms slightly decreased, as shown in Figure 4e aligning with the least compressible c- and b-axes. Consequently, the overlapping area between the phenyl and MR framework moieties considerably expanded (Figure 4d). The interplanar distance Dth (Figure 4c) also decreased from 1 atm to 9.0 GPa. Notably, the Dth distance exceeded 3.70 Å at pressures below 3.00 GPa (Figure 4e), which surpassed the range of typical π‒π interaction distance. This result, together with the small overlapping area between the π-conjugated moieties (Figure 4d), indicated that BNC adopted isolated and discrete π–π dimer stacking at pressures below 3.5 GPa. Upon further compression, the packing of the molecules became denser such that the π‒π dimers changed to a non-discrete state. Therefore, above 4 GPa, the PL quenching of the BNC crystals was attributed to the intensified nonradiative relaxation caused by long-range excimers.[35,39] Specifically, the non-discrete stacking of these dimers, caused by the high pressure (>4 GPa), resulted in a low PLQY of excimer fluorescence. The isolated dimer of BNC was extracted from the corresponding unit cell at each pressure for subsequent analysis. As illustrated in Figure 4b, the interplanar distance Dtt within the isolated dimers consistently decreased with increasing pressure. Interestingly, at pressures up to 0.50 GPa, the distance Dtt decreased to as small as 3.30 Å, which was rarely observed at ambient pressure. This distance was slightly smaller than the π–π interplanar distance of the excimer equilibrium geometry (Re = 3.330 Å).[40] Thus, the dimer of BNC likely exhibited a “compressed” ground state characterized by a shortened π–π distance and increased π–π overlap (Vide supra).
[IMAGE OMITTED. SEE PDF]
Piezochromic Mechanism of BNC
The Independent gradient model (IGM) analysis is a valuable tool for visualizing intermolecular interactions in three dimensions (3D). It was utilized to monitor the evolution of these interactions under varying pressures. In the isosurface representation, the blue (red) color indicates the strength of attractive interactions (steric effects), while the green zone signifies π–π interactions. In our case, as depicted in Figure S14 (Supporting Information), the green areas gradually expanded with increasing pressure, demonstrating the presence of strong π–π interactions and π–π overlap within the isolated dimers. Moreover, the pressure-dependent average PL lifetime τ is calculated by τ = (A1τ12 + A2τ22)/(A1τ1 + A2τ2). During compression, the PL lifetime of the shoulder peak at 587 nm significantly increased below 0.62 GPa, as shown in Figure 5a. This increase coincided with the PL quenching observed at pressures below 0.48 GPa (Figure 3e). This result further verified that the π–π distance (Dtt) decreased and the π–π overlap increased within the isolated dimers; thus, a “compressed” ground state was formed. Above 1 GPa, the increase in the PL lifetime leveled off. Then, we evaluated the roles of the main excited-state energy dissipation ways in exciton‒vibration couplings. The strength of the exciton‒vibration couplings was measured by the reorganization energy (λes, S1→S0). Upon compression, from 0 to 3.0 GPa, the λes of BNC in crystalline aggregates monotonically decreased, as depicted in Figure 5b, thus effectively restricting the exciton‒vibration coupling of BNC. With a compression of more than 3 GPa, the value of λes remained close to 680 cm−1, suggesting that the exciton‒vibration coupling was not further weakened. In low-pressure region, the reduction in λes was primarily attributed to the low-frequency vibrations (ω < 200 cm−1, inset in Figure 5b) caused by the rotation of the phenyl rings. To further investigate the molecular vibrations, in situ infrared (IR) spectra ranging from 800–3300 cm−1 were recorded at high pressures. As depicted in Figure 5c, the relatively prominent peaks at 1528 and 1603 cm−1 in the IR spectra at ambient pressure were attributed to the C═C bond stretching vibration of the aromatic ring. With increasing pressure, these peaks representing the C═C bond stretching vibration showed minimal changes in absorption shape and peak position. This observation verified that the decrease in λes for BNC was not influenced by the C═C bond stretching vibration under high pressure. The distinct peaks at 2963.4 and 2603.2 cm−1 corresponded to the C─H and B─H bond stretching vibrations of the BNC, respectively. Impressively, the intensity of C─H and B─H stretching vibration modes decreased significantly, while their peak widths increased below 3.0 GPa (Figure 5c,d). These observations suggest that the distances between the C/B atoms and the H atoms of adjacent molecules decreased under high pressure, leading to stronger molecular interactions. Consequently, these interactions restricted the B─H and C─H stretching vibrations, thereby inhibiting the non-radiative vibrational processes described by Equation (2). However, above 3.0 GPa, the rate of the above change began to slow (Figure 5d), indicating a slight weakening of the exciton‒vibration coupling.
[IMAGE OMITTED. SEE PDF]
According to the experimental evidence presented, the thorough analysis of the experimental and calculated results led to some reasonable deductions regarding the complicated changes in the PL spectra of the BNC crystals under high pressure as follows: 1) below 0.48 GPa, the pressure caused the contraction of the π‒π distance Dtt and an increase in the π‒π overlap ratio within dimers to facilitate the enhanced excimer and polarization effects; consequently, Eb experienced a slight decrease, leading to a reduction in the PL intensity (Figure 3a); and 2) from 0.48 to 3.0 GPa, the BNC molecules exhibited discrete π‒π dimer stacking; the molecules stacked more tightly within the isolated dimers, resulting in the formation of a “compressed” ground state. Thus, the B─H and C─H stretching vibrations, as well as low-frequency rotation, were suppressed by high pressure, weakening the exciton‒vibration coupling. This suppression ultimately led to an increase in the PL intensity and spectral narrowing of the BNC crystal (Figure 3b). Upon further compression, the stacking of these dimers produced a non-discrete state, where long-range excimers occurred, and PL quenching occurred (Figure 3c).
Conclusion
In summary, we investigated the piezochromic behavior of a classical MR material and correlated the alterations in intermolecular interactions and molecular vibrations with the photophysical properties. Interestingly, an abnormal phenomenon of pressure-induced emission enhancement and spectral narrowing was observed in the BNC crystal when the pressure exceeded 0.48 GPa and reached a maximum at 3.0 GPa. Joint in situ experimental and theoretical analyses revealed that the unusual piezochromism was associated with a change in the exciton binding energy and exciton–vibrational coupling. Below 0.48 GPa, the reduction of the Eb had a greater effect on the relaxation decay of the excited state than the suppression of the exciton–vibrational coupling. Thus, the intensity of PL spectra slightly decreased upon compression. Between approximately 0.48 and 3.0 GPa, the distance Dtt of the dimers was less than 3.3 Å. Thus, a dimer could be considered a “compressed” emitting unit within this pressure range. The stretching vibrations of B─H and C─H, along with the low-frequency rotation, were suppressed within the isolated dimer, thereby weakening the exciton‒vibration coupling. Moreover, the Dth distance exceeded 3.7 Å, indicating that these dimers were isolated and independent and leading to a minor alteration in the polarization effects. the Eb value of the BNC crystal likely underwent a slight change. Consequently, in the process, the weakening of the exciton–vibrational couplings was dominant; this ultimately resulted in a clear increase in the PL intensity and a slight narrowing of the PL spectra. This study provides an unconventional strategy to better understand the relationships among the intermolecular interactions, molecular vibrations, and luminescence behavior; this understanding is needed for the design of high-performance narrowband organic light-emitting materials, according to the desired application.
Experimental Section
Preparation of BNC
Both the commercially available B/N skeleton boric acid ester (BCpin) and 1-(4-bromophenyl)-o-carborane were purified through a recrystallization process. BNC were readily prepared via the Suzuki coupling reaction of BCpin with 1-(4-bromophenyl)-o-carborane (refer to the Supporting Information for the detailed synthetic procedure). The chemical structures of the MR emitters were thoroughly characterized using nuclear magnetic resonance (NMR) spectroscopy (Figures S1 and S2, Supporting Information), mass spectrometry (MS) (Figure S3, Supporting Information), and single-crystal X-ray diffraction.
Acknowledgements
Z.L. and Q.Z. contributed equally to this work. This work was supported by the National Nature Science Foundation of China (52073089, 52273175, and 22203029) (Y.Z., C.L., and Q.Z.), the Natural Science Foundation of Zhejiang Province (LQ24B020001) (F.S.), the Special Construction Project Fund for Shandong Province Taishan Scholars, the Huzhou Key Laboratory of Environmental Functional Materials and Pollution Control.
Conflict of Interest
The authors declare no conflict of interest.
Data Availability Statement
The data that support the findings of this study are available on request from the corresponding author. The data are not publicly available due to privacy or ethical restrictions.
T. Fleetham, G. Li, L. Wen, J. Li, Adv. Mater. 2014, 26, 7116.
Y. Kondo, K. Yoshiura, S. Kitera, H. Nishi, S. Oda, H. Gotoh, Y. Sasada, M. Yanai, T. Hatakeyama, Nat. Photonics 2019, 13, 678.
M. Yang, I. S. Park, T. Yasuda, J. Am. Chem. Soc. 2020, 142, [eLocator: 19468].
S. Oda, T. Sugitani, H. Tanaka, K. Tabata, R. Kawasumi, T. Hatakeyama, Adv. Mater. 2022, 34, [eLocator: 2201778].
H. Uoyama, K. Goushi, K. Shizu, H. Nomura, C. Adachi, Nature 2012, 492, 234.
P. Rajamalli, N. Senthilkumar, P. Gandeepan, P. Huang, M. Huang, C. Ren‐Wu, C. Yang, M. Chiu, L. Chu, H. Lin, C. Cheng, J. Am. Chem. Soc. 2016, 138, 628.
P. Data, P. Pander, M. Okazaki, Y. Takeda, S. Minakata, A. P. Monkman, Angew. Chem., Int. Ed. 2016, 55, 5739.
T. Hatakeyama, K. Shiren, K. Nakajima, S. Nomura, S. Nakatsuka, K. Kinoshita, J. Ni, Y. Ono, T. Ikuta, Adv. Mater. 2016, 28, 2777.
Y. Zhang, J. Wei, L. Wang, T. Huang, G. Meng, X. Wang, X. Zeng, M. Du, T. Fan, C. Yin, D. Zhang, L. Duan, Adv. Mater. 2023, 35, [eLocator: 2209396].
S. Luo, J. Wang, N. Li, X. Song, X. Wan, K. Li, C. Yang, Angew. Chem., Int. Ed. 2023, 62, [eLocator: e202310943].
Y. Yu, Z. Feng, X. Meng, L. Chen, F. Liu, S. Yang, D. Zhou, L. Liao, Z. Jiang, Angew. Chem., Int. Ed. 2023, 62, [eLocator: e202310047].
Y. Wu, X. Liu, J. Liu, G. Yang, S. Han, D. Yang, X. Cao, D. Ma, Z. Bin, J. You, Mater. Horiz. 2023, 10, 3785.
S. Jiang, Y. Yu, D. Li, Z. Chen, Y. He, M. Li, G. Yang, W. Qiu, Z. Yang, Y. Gan, J. Lin, Y. Ma, S. Su, Angew. Chem., Int. Ed. 2023, 62, [eLocator: e202218892].
G. Meng, H. Dai, Q. Wang, J. Zhou, T. Fan, X. Zeng, X. Wang, Y. Zhang, D. Yang, D. Ma, D. Zhang, L. Duan, Nat. Commun. 2023, 14, 2394.
P. Jiang, J. Miao, X. Cao, H. Xia, K. Pan, T. Hua, X. Lv, Z. Huang, Y. Zou, C. Yang, Adv. Mater. 2022, 34, [eLocator: 2106954].
Y. Zhang, J. Wei, D. Zhang, C. Yin, G. Li, Z. Liu, X. Jia, J. Qiao, L. Duan, Angew. Chem., Int. Ed. 2022, 61, [eLocator: e202113206].
H. Kim, H. Cheon, D. Lee, W. Lee, J. Kim, Y. Kim, S. Yoo, Sci. Adv. 2023, 9, [eLocator: eadf1388].
K. Stavrou, A. Danos, T. Hama, T. Hatakeyama, A. Monkman, ACS Appl. Mater. Interfaces 2021, 13, 8643.
W. Stadler, D. M. Hofmann, H. C. Alt, T. Muschik, B. K. Meyer, E. Weigel, G. Müller‐Vogt, M. Salk, E. Rupp, K. W. Benz, Phys. Rev. B 1995, 51, [eLocator: 10619].
A. Nitzan, S. Mukamel, J. Jortner, J. Chem. Phys. 1975, 63, 200.
I. R. Gould, D. Noukakis, L. Gomez‐Jahn, R. H. Young, J. L. Goodman, S. Farid, Chem. Phys. 1993, 176, 439.
S. Tong, J. Dai, J. Sun, Y. Liu, X. Ma, Z. Liu, T. Ma, J. Tan, Z. Yao, S. Wang, H. Zheng, K. Wang, F. Hong, X. Yu, C. Gao, X. Gu, Nat. Commun. 2022, 13, 5234.
Y. Shen, Z. An, H. Liu, B. Yang, Y. Zhang, Angew. Chem., Int. Ed. 2023, 62, [eLocator: e202214483].
G. Liu, L. Kong, J. Gong, W. Yang, H. Mao, Q. Hu, Z. Liu, R. Schaller, D. Zhang, T. Xu, Adv. Funct. Mater. 2017, 27, [eLocator: 1604208].
H. Liu, Y. Gu, Y. Dai, K. Wang, S. Zhang, G. Chen, B. Zou, B. Yang, J. Am. Chem. Soc. 2020, 142, 1153.
H. Luo, S. Guo, Y. Zhang, K. Bu, H. Lin, Y. Wang, Y. Yin, D. Zhang, S. Jin, W. Zhang, W. Yang, B. Ma, X. Lu, Adv. Sci. 2021, 8, [eLocator: 2100786].
Y. Xia, C. Zhu, F. Cao, Y. Shen, M. Ouyang, Y. Zhang, Angew. Chem., Int. Ed. 2023, 62, [eLocator: e202217547].
Y. Wang, J. Yang, M. Fang, Y. Yu, B. Zou, L. Wang, Y. Tian, J. Cheng, B. Tang, Z. Li, Matter 2020, 3, 449.
C. Zhai, X. Yin, S. Niu, M. Yao, S. Hu, J. Dong, Y. Shang, Z. Wang, Q. Li, B. Sundqvist, B. Liu, Nat. Commun. 2021, 12, 4084.
Z. Yang, Z. Fu, H. Liu, M. Wu, N. li, K. Wang, S. Zhang, B. Zou, B. Yang, Chem. Sci. 2023, 14, 2640.
S. Lu, G. Xiao, L. Sui, T. Feng, X. Yong, S. Zhu, B. Li, Z. Liu, B. Zou, M. Jin, J. Tse, H. Yan, B. Yang, Angew. Chem., Int. Ed. 2017, 56, 6187.
Y. Xu, C. Li, Z. Li, Q. Wang, X. Cai, J. Wei, Y. Wang, Angew. Chem., Int. Ed. 2020, 59, [eLocator: 17442].
G. Chen, J. Wang, W. Chen, Y. Gong, N. Zhuang, H. Liang, L. Xing, Y. Liu, S. Ji, H. Zhang, Z. Zhao, Y. Huo, B. Tang, Adv. Funct. Mater. 2023, 33, [eLocator: 2211893].
H. Liu, L. Yao, B. Li, X. Chen, Y. Gao, S. Zhang, W. Li, P. Lu, B. Yang, Y. Ma, Chem. Commun. 2016, 52, 7356.
Q. Luo, L. Li, H. Ma, C. Lv, X. Jiang, X. Gu, Z. An, B. Zou, C. Zhang, Y. Zhang, Chem. Sci. 2020, 11, 6020.
Y. Zhu, D. He, C. Wang, X. Han, Z. Liu, K. Wang, J. Zhang, X. Shen, J. Li, Y. Lin, C. Wang, Y. He, F. Zhao, Angew. Chem., Int. Ed. 2024, 63, [eLocator: e202316227].
X. Dai, Z. Zhang, Y. Jin, Y. Niu, H. Cao, X. Liang, L. Chen, J. Wang, X. Peng, Nature 2014, 515, 96.
L. Zhu, J. Zhang, Y. Guo, C. Yang, Y. Yi, Z. Wei, Angew. Chem., Int. Ed. 2021, 60, [eLocator: 15348].
K. Nagura, S. Saito, H. Yusa, H. Yamawaki, H. Fujihisa, H. Sato, Y. Shimoikeda, S. Yamaguchi, J. Am. Chem. Soc. 2013, 135, [eLocator: 10322]
H. Liu, Y. Dai, Y. Gao, H. Gao, L. Yao, S. Zhang, Z. Xie, K. Wang, B. Zou, B. Yang, Y. Ma, Adv. Opt. Mater. 2018, 6, [eLocator: 1800085].
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2025. This work is published under http://creativecommons.org/licenses/by/4.0/ (the "License"). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
Polycyclic multiple resonance (MR) molecules reveal narrowband emission, making them very promising emitters for high color purity display. Nevertheless, they still have challenges such as aggregation‐induced emission quenching and spectral broadening. Overcoming these obstacles requires an in‐depth understanding of the correlations among the alterations in their geometries, packing structures, and molecular vibrations and their corresponding changes in their photoluminescence (PL) properties. Herein, it is demonstrated that high‐pressure infrared, UV−visible absorption, and fluorescence spectroscopies can be combined with computational results to elucidate the influence of the subtle structural variations on the exciton‒vibration couplings and their PL properties. An ortho‐carborane‐decorated MR emitter (BNC) is a piezochromic molecule and exhibits emission enhancement under high pressure. A thorough analysis of the in situ experimental measurements and calculated results reveals that the pressure‐induced changes in the exciton binding energy and exciton‒vibration couplings are responsible for the unusual piezochromism. This research provides insights into the structure‒fluorescence relationship and potential for high‐pressure techniques to optimize MR materials for advanced organic light‐emitting diodes (OLEDs) applications.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Details


1 Key Laboratory of the Ministry of Education for Advanced Catalysis Materials, Department of Chemistry, Zhejiang Normal University, Jinhua, P. R. China
2 Department of Materials Chemistry, Huzhou University, Huzhou, P. R. China
3 School of Physics Science and Information Technology, Liaocheng University, Liaocheng, P. R. China