Open Access
2 February 2018 Solvent-dependent investigation of carbazole benzonitrile derivatives: does the 3LE−1CT energy gap facilitate thermally activated delayed fluorescence?
Takuya Hosokai, Hiroki Noda, Hajime Nakanotani, Takanori Nawata, Yasuo Nakayama, Hiroyuki Matsuzaki, Chihaya Adachi
Author Affiliations +
Abstract
The photophysical properties of six types of carbazole benzonitrile (CzBN) derivatives are investigated in different solvents to examine the thermally activated delayed fluorescence (TADF) activation via reducing the energy gap between the singlet charge-transfer and triplet locally excited states, ΔEST(LE). Relative to the ΔEST(LE) values for the CzBN derivatives in the low polarity solvent toluene (ϵ∼2), a reduction of ΔEST(LE) for the CzBN derivatives in the polar solvent acetonitrile (ϵ∼37) was confirmed while maintaining fairly constant ΔEST values. Notably, TADF activation was observed in acetonitrile for some CzBN derivatives that are TADF inactive in toluene. A numerical analysis of various rate constants revealed the cause of TADF activation as an increase in the reverse intersystem crossing rate and a suppression of the non-radiative decay rate of the triplet states. The positive effect of ΔEST(LE) was limited, however, as an excessive decrease in ΔEST(LE) facilitates the nonradiative deactivation of the triplet states, leading to a loss of the TADF efficiency. This paper shows that ΔEST(LE) provides a measure of TADF activation and that appropriate regulation of ΔEST(LE) is required to achieve high TADF efficiency.

1.

Introduction

Thermally activated delayed fluorescence (TADF) of organic molecules has attracted tremendous attention owing to its potential application to organic light-emitting diodes (OLEDs) and has also renewed interest in photochemistry and photophysics.1,2 Uoyama et al.3 established a design concept for the synthesis of highly efficient TADF molecules by connecting electron-donating and -accepting substituents with a large torsion angle. Such a structure can separate the wavefunction overlap between the highest occupied molecular orbital and the lowest unoccupied molecular orbital, leading to a significant reduction in the energy gap (ΔEST) between the first excited singlet state (S1) and the first excited triplet state (T1). Notably, the ΔEST diminishes to thermal energy at room temperature, 25  meV, and typically <100  meV. Thus, this approach affords a thermal pathway to convert from the T1 state to the S1 state via reverse intersystem crossing (RISC), thereby facilitating TADF. After Uoyama’s report,3 a number of TADF molecules were rapidly synthesized and exhibited a high internal photoluminescence quantum yield (PLQY) of TADF (ϕDF) in various emission colors, including blue.410 Additionally, the device performance of OLEDs, such as external electroluminescence quantum efficiency and operational stability, is being improved by this approach.1114

In addition to these exciting applications of TADF, the fundamental processes in TADF are also attracting considerable interest in both experimental and theoretical fields.1526 In this case, the goal is to achieve a sufficient understanding of these processes to realize a high RISC rate (kRISC) by judicious designing and optimizing the most suitable chemical structures, emission environments, and so on. Increasing the kRISC suppresses triplet state quenching and is, therefore, beneficial for reducing roll-off in OLEDs and organic laser diodes.2729

One important question of the TADF mechanism concerns the rate of the various electron transfer processes. That is, although ΔEST values at thermal energy or even less have been attained, and intersystem crossing rates (kISC) as high as 108  s1 have been obtained, the corresponding kRISC values were often limited to a maximum of 106  s1. Moreover, ϕDF values close to 80% were achieved, even with a relatively appreciable ΔEST value of 0.2  eV.24 Theoretical simulations have indicated that selecting certain molecular orbitals between the S1 and T1 states is essential to achieve a high kRISC; known as El-Sayed’s rule, it has been shown that high kRISC values can be achieved by exploiting the singlet charge-transfer (CT1) and triplet locally excited (LE3) states.18,19,21,25 In addition, some groups have proposed that second-order spin-orbit coupling enables utilization of higher order triplet states, which further facilitates RISC.18,20,23 Conventional and time-dependent density functional theory calculations based on such orbital selection rules and/or second-order perturbation theory have succeeded in reproducing the experimentally determined kRISC values of some, but not all, of the organic molecules investigated.21 Thus, a broad consensus of the TADF mechanism has yet to be reached.

Another aspect of the selection rules for RISC is our recent proposal based on experimental data that the energy gap between CT1 and LE3, so-called ΔEST(LE), is an important parameter to signify TADF [Fig. 1(a)].24 The carbazole benzonitrile (CzBN) derivatives 2CzBN, o-3CzBN, m-3CzBN, p-3CzBN, 4CzBN, and 5CzBN [Fig. 1(b)] have shown relatively similar ΔEST values of 0.2  eV, as determined by their corresponding fluorescence and phosphorescence spectra, together with fairly different kRISC values from zero to 105  s1.24 Assuming that their LE3 values are similar, which arise from structurally common Cz moieties (vide infra), the ΔEST(LE) values estimated from the LE3(Cz) and CT1 levels of each CzBN derivative varied from 0.06 to 0.23 eV because of different CT1 levels for each CzBN derivative. As a result, ΔEST(LE) and kRISC, but not ΔEST, were found to be correlated; specifically, decreasing values of ΔEST(LE) lead to an increase in kRISC. For this reason, ΔEST(LE) was proposed as an additional measure, in conjunction with ΔEST, for determining whether molecules could exhibit TADF. However, further investigations are needed to verify the generality of this concept; in particular, it is desirable to establish whether tuning ΔEST(LE) can indeed facilitate TADF and vice versa.

Fig. 1

(a) Definition of ΔEST and ΔEST(LE). (b) Chemical structures of CzBN derivatives.

JPE_8_3_032102_f001.png

In this paper, we investigated the solvent dependence of the photophysical properties of six types of CzBN derivatives. It is known that a host matrix can tune CT state energies and that large environmental dielectric fields stabilize CT states by orienting the polarization of the surrounding solvent. Notably, LE states are hardly affected by this phenomenon.30 This means that selecting the type of host matrix (e.g., solvent) can modulate the energy levels of CzBN derivatives and give rise to different photophysical properties. We previously investigated the photophysical properties of CzBN derivatives in toluene, whose dielectric constant (ϵ) is 2 at room temperature. In the present study, the strong polar solvent acetonitrile (ϵ37) was used to investigate the photophysical properties of the same six CzBN derivatives, the results of which are compared with the previous results determined in toluene.24 The large ϵ of acetonitrile is expected to stabilize the CT1 and CT3 states and, thus, tune the energy levels, in particular ΔEST(LE), of the CzBN derivatives.

2.

Experimental

All the materials were synthesized according to Ref. 24 and purified by thermal sublimation. Solutions were prepared by dissolving the purified molecules in toluene or acetonitrile, both of which were of spectrochemical analysis grade (purity: 99.8%, Wako Pure Chemical Industries, Ltd.). The solution concentration ranged from 104 to 105  M depending on the samples.

Optical properties of the solutions were characterized by measuring the steady-state ultraviolet-visible (UV-vis) absorption/PL spectra and transient PL decay curves (hereafter called transient PL). UV-vis absorption spectra were acquired using a UV-VIS-NIR scanning spectrophotometer (UV-3100PC, Shimazu), whereas PL spectra were recorded using a fluorescence spectrophotometer (LS50B, Perkin Instruments or Lambda 950-PKA Instrument, Perkin–Elmer). Transient PLs were measured using a fluorescence lifetime spectrometer (C11367-01, Hamamatsu Photonics or a FluoroCube, Horiba). The PLQY of the solutions was measured using an absolute PLQY measurement system (C11347-01, Hamamatsu Photonics) with an excitation wavelength of 337 nm under two conditions: as prepared and after degassing by dry argon gas to eliminate triplet state deactivation. Before the transient PL measurements, the solutions were also deoxygenated with dry nitrogen or argon gas for the same reason. The triplet state lifetime of 2CzBN in acetonitrile was determined by microsecond-transient absorption spectroscopy using the third harmonic of fundamental light (1064 nm) of a Nd3+:YAG laser as the pump light (wavelength, 355 nm; FWHM pulse, <150  ps; repetition, 10 Hz) and a xenon steady-state lamp as the probe light. Full details of this design are described in Ref. 31. The irradiated intensity of the pump laser was set to 0.7  mJ/cm2.

3.

Results and Discussion

Figures 2(a)2(f) show the UV–vis absorption, fluorescence, and phosphorescence spectra of the six CzBN derivatives in toluene (upper) and acetonitrile (bottom). In the absorption spectra, ππ* transitions at a single Cz moiety [Cz(ππ*)] are commonly observed in the UV region (3.7 to 4.0 eV). Below this region, two types of CT transitions are observed depending on the molecules: CT1, which originates from a CT transition from one Cz moiety to the center BN moiety, located at 3.4 to 3.6 eV, and CT2, which is a CT transition among two Cz moieties at para-positions [2 and 5 or 3 and 6, see blue-colored Czs in Fig. 1(b)] and the BN moiety at the lowest energy region (3.0 to 3.2 eV).24 Both CT transitions are apparent in the absorption spectrum of 5CzBN [Fig. 2(f)]. Their formation in 5CzBN can be understood in terms of its chemical structure, which is composed of two para-Cz pairs and a single Cz moiety opposite the cyano group; collectively, these structural features fulfill the electronic condition of both CT transitions. By comparison, only the CT1 transition was present in the spectra of 2CzBN, o-3CzBN, and m-3CzBN, and only the CT2 transition was evident in the spectra of p-3CzBN and 4CzBN. In terms of their molecular structures, the former three have no such Cz pairs, and the latter two possess Cz pairs. The structure-dependent absorption of the CzBN derivatives in toluene was also observed viz. their corresponding transient absorption spectra, as reported in Ref. 24.

Fig. 2

UV–vis absorption, fluorescence, and phosphorescence (77 K) spectra of (a) 2CzBN, (b) o-3CzBN, (c) m-3CzBN, (d) p-3CzBN, (e) 4CzBN, and (f) 5CzBN in toluene (upper) and acetonitrile (bottom). Threshold (or partial peak top) energies in the fluorescence and phosphorescence spectra are depicted by blue or red vertical arrows, respectively.

JPE_8_3_032102_f002.png

The three distinct absorption features in Figs. 2(a)2(f) [i.e., Cz (ππ*), CT1, and CT2] indicate different photophysical behavior of the CzBN derivatives in toluene and acetonitrile. The peak positions of the Cz (ππ*) transition remain almost unchanged, whereas both the CT1 and CT2 transitions exhibit a blue shift and a decrease in intensity in acetonitrile. Moreover, the intensity decrease is much stronger for CT1 compared with CT2; for instance, the peak structure of the CT1 transition changes to a shoulder-like feature in acetonitrile for 5CzBN. These results suggest that both the CT1 and CT2 transitions contain nπ* character, but the solvent effect differs between local CT (CT1) and delocalized (at para-Cz pairs) CT (CT2), presumably due to the distinct electron distributions in their excited states.

Solvent effects were also evident in the fluorescence and phosphorescence spectra of the six CzBN derivatives. Essentially, both the fluorescence and phosphorescence spectra show a red shift in acetonitrile, as seen in Figs. 2(a)2(f), which we attribute to stabilization of both the CT1 and CT3 states. Only the phosphorescence spectrum of 2CzBN in toluene exhibits distinctly sharp vibronic progressions, which are attributable to emission from LE3 in the Cz moieties.24 This result also suggests that the LE3 state of 2CzBN in toluene is lower in energy than the CT3 state. By comparison, the phosphorescence spectrum of 2CzBN in acetonitrile is rather broadened, suggesting an exchange of the T1 from LE3 to CT3 or an increase of electronic coupling between the two states, i.e., the strengthening of state mixing, due to their energy proximity because of the CT3 stabilization.32,33

On the basis of the fluorescence and phosphorescence results, the ΔEST and ΔEST(LE) values of each CzBN derivative were estimated and are given in Table 1, where the S1 and T1 energies used for the estimation are also depicted. For the ΔEST(LE), the energy of the LE3 state in the phosphorescence spectrum of 2CzBN was used, as with our previous report.24 Here we note that both the ΔEST and ΔEST(LE) values in the toluene samples are slightly differed from the previous reports owing to the use of different apparatus and measurement conditions. As seen in Table 1, the ΔEST values in toluene range from 0.16 to 0.27 eV, and these values are not appreciably reduced in acetonitrile, except for 5CzBN; in this case, 5CzBN showed a decrease in ΔEST from 0.16 to 0.07 eV. In this context, the kRISC value of only 5CzBN is anticipated to increase according to the relation between ΔEST and kRISC: kRISC1/exp(ΔEST/kT), where k is the Boltzmann constant and T is the temperature. However, as will be shown, this expectation is misleading because kRISC is actually influenced significantly by ΔEST(LE). As expected from the decrease in the CT3 energies, the ΔEST(LE) values of all the CzBN derivatives decreased in acetonitrile. In particular, the ΔEST(LE) values of o-3CzBN and m-3CzBN are reduced largely from 0.16 to 0.04 eV and from 0.20 to 0.09 eV, respectively. These values are less than that of p-3CzBN examined previously, which was the only isomer of the three that showed TADF in toluene.24 On the basis of the present results, TADF is expected to be active for o-3CzBN and m-3CzBN in acetonitrile.

Table 1

ΔEST and ΔEST(LE) of CzBN derivatives in toluene and acetonitrile.

2CzBNo-3CzBNm-3CzBNp-3CzBN4CzBN5CzBN
ΔEST (eV)a,cToluene0.260.180.270.240.220.16
Acetonitrile0.310.190.290.310.250.07
ΔEST(LE) (eV)b,cToluene0.260.160.200.140.010.08
Acetonitrile0.210.040.090.100.040.22
S1 level (eV)Toluene3.273.173.213.153.022.93
Acetonitrile3.223.053.103.112.972.79
T1 level (eV)Toluene3.012.992.952.922.802.78
Acetonitrile2.912.862.802.802.722.72

aEnergy gap calculated from the S1 and T1 energy levels estimated from the threshold of the fluorescence spectra and peak (2CzBN in toluene) or threshold (2CzBN in acetonitrile and other CzBN derivatives) of the phosphorescence spectra, respectively [Figs. 2(a)–2(f)].

bEnergy gap calculated from the CT1 (S1) and LE3 (T1 of 2CzBN in toluene), respectively.

cThe error bar of ΔEST and ΔEST(LE) is a maximum of ±0.05  eV.

To examine the ΔEST(LE) values in the context of TADF properties, we next discuss the results of transient PL for all the sample solutions. As seen in Figs. 3(a)3(f), the transient PL measured for toluene (red line) are indicated in the bottom axes as a function of their delay time, whereas those of acetonitrile (blue line) are indicated in the upper axes. Note that the time scale for the three 3CzBN isomers differs by orders of magnitude in toluene (μs) and acetonitrile (ms). As expected from the results of ΔEST(LE), o-3CzBN and m-3CzBN, which are TADF inactive in toluene, become TADF active in acetonitrile. Delayed PL components with time constants (τDF) of 66 and 220  μs for o-3CzBN [Fig. 3(b)] and m-3CzBN [Fig. 3(c)], respectively, were confirmed. These τDF values are comparable with that of p-3CzBN in acetonitrile [103  μs, Fig. 3(d)]. Importantly, ϕDF of the three 3CzBN isomers are similar: 21% for o-3CzBN, 18% for m-3CzBN, and 24% for p-3CzBN. These values were calculated by subtracting the PLQY in air (ϕPL,air) from the PLQY measured after argon bubbling (ϕPL,degassed). After analyzing the photophysical properties of all the solution samples according to Ref. 27, the values of kRISC are obtained as 3.0×104  s1 for o-3CzBN, 6.4×103  s1 for m-3CzBN, and 2.1×104  s1 for p-3CzBN. The kRISC of p-3CzBN is slightly improved from 1.3×104  s1 in toluene, but that of o-3CzBN and m-3CzBN improved at least 1 to 2 orders of magnitude, based on the assumption that 102  s1 is a lower limit of our estimation. These results clearly indicate that a smaller ΔEST(LE) increases the value of kRISC, thereby facilitating TADF. Therefore, TADF can be triggered from being inactive to active by environmental effects. Note that the ΔEST(LE) value of 0.20 eV for 2CzBN in acetonitrile is still too large for TADF to develop.

Fig. 3

Comparison of transient PL of CzBN derivatives in toluene (red line, bottom axis) and acetonitrile (blue line, upper axis): (a) 2CzBN, (b) o-3CzBN, (c) m-3CzBN, (d) p-3CzBN, (e) 4CzBN, and (f) 5CzBN.

JPE_8_3_032102_f003.png

It seems that the effect of reducing ΔEST(LE) on kRISC is limited, as the kRISC values of 4CzBN and 5CzBN improved only slightly from 1.8×105  s1 to 2.1×105  s1 and from 2.4×105  s1 to 2.6×105  s1 in acetonitrile, respectively. This observation may be deduced from the fact that their ΔEST(LE) values become more negative, which means that the LE3 state lies farther from the CT1 state in acetonitrile. This negative shift in ΔEST(LE) does not impact the ϕDF value of 4CzBN, although it does significantly decrease the ϕDF value of 5CzBN from 76% in toluene to 24% in acetonitrile. This decrease is caused by an increase in the nonradiative decay rate from T1 to the ground state (S0) (knr,T) from 4.2×103  s1 (in toluene) to 9.3×104  s1 (in acetonitrile). As mentioned above, however, kRISC does not decrease in acetonitrile, indicating that coupling between the LE3 and CT1 states is important for RISC.

Additionally, it is instructive to discuss the impact of the solvent effect for other rate constants relevant to the electron transfer processes in the six CzBN molecules examined. In acetonitrile, the radiation decay rate from S1 to S0 (kr,S) and kISC tends to decrease (Table 2). However, this is not the cause of the increase in the nonradiative decay rate for the S1 to S0 transition (knr,S), since the ϕPL,air of the CzBN derivatives, which assumes a PLQY of prompt fluorescence, remains unchanged or even increases, except for 2CzBN and o-3CzBN. Although knr,T of 5CzBN is increased in acetonitrile, the other CzBN derivatives showed a decrease of knr,T. Overall, it is clear that ϕPL,degassed of the three 3CzBN isomers and 4CzBN increased in acetonitrile.

Table 2

Photophysical properties of the CzBN derivatives.

ϕPL,air/ϕPL,degassed (%)ϕDF (%)τpr/τDF(ns)/(μs)kr,S (106  s−1)knr,T (103  s−1)kISC (107  s−1)kRISC (103  s−1)
2CzBNToluene15/23010.9/1.3a13.8783a7.80
Acetonitrile9/20012.4/47.5a7.321.1a7.30
o-3CzBNToluene21/31018.5/15a11.465a4.30
Acetonitrile12/332114.3/66.28.411.56.230.0
m-3CzBNToluene15/1703.6/39a41.726a23.60
Acetonitrile15/331810.1/220.314.93.68.46.4
p-3CzBNToluene10/1441.2/3583.32775.012.7
Acetonitrile13/37249.6/102.713.67.09.120.6
4CzBNToluene9/62531.6/3656.31756.9179.8
Acetonitrile9/73645.4/36.916.68.016.8212.0
5CzBNToluene9/85763.8/3923.74.223.9237.9
Acetonitrile14/382416.2/7.88.693.05.3257.0

aVirtual values as determined by measuring the triplet-state absorption decay curve in microsecond-TAS (see Ref. 24).

The change in the rate constants in acetonitrile is qualitatively similar to 4CzIPN,34 implying that solvent effects are a general phenomenon for TADF molecules. The only difference is that the TADF efficiency in toluene, the most nonpolar solvent used by Ishimatsu et al.,34 demonstrates the best value, suggesting that the absolute value of ΔEST(LE) is a minimum in toluene for 4CzIPN, as for 5CzBN. In fact, the LE3 state of 4CzIPN is already above the CT1 and CT3 states (Fig. 4). This result further supports the solvent effect for TADF efficiency and suggests that the most appropriate “solvent environment”, that will couple the relevant electronic states, CT1, CT3, and LE3, for TADF efficiency depends on a combination of both the emitter molecules and the solvent.

Fig. 4

Energy level diagram of 4CzIPN in toluene. The S1 and T1 energies and the ΔEST value are taken from Ref. 24. The LE3 state at IPN is reported in Ref. 35.

JPE_8_3_032102_f004.png

4.

Conclusion

This study examined the solvent dependence of the photophysical properties of a series of CzBN derivatives. Compared with toluene (weakly polar), acetonitrile (strongly polar), which lowers the ΔEST(LE), facilitates TADF for o-3CzBN and m-3CzBN and improves the TADF efficiency for p-3CzBN and 4CzBN. Kinetically, these results can be explained by both an increase in kRISC and a decrease in knr,T. However, a strongly polar solvent is not always beneficial for TADF efficiency, as 5CzBN (and also 4CzIPN in Ref. 34) showed a larger decrease in TADF efficiency due to the consequent increase in knr,T. As a result, the polar host environment has an advantage for kRISC, but the appropriate polarity is required for respective molecules to maximize the effect in TADF efficiency. Overall, since close proximity of the LE3 and CT1 states facilitates TADF, its efficiency can be improved by tuning both the ΔEST and ΔEST(LE), which are realized by a combination of (1) the design of the donor-accepter geometry (for CT1 and CT3) and (2) the selection of donor and accepter moieties (for LE3).

Acknowledgments

This work was supported, in part, by the International Institute for Carbon Neutral Energy Research (WPI-I2CNER) sponsored by the Ministry of Education, Culture, Sports, Science and Technology (MEXT), and JSPS-KAKENHI Grant Nos. JP16K14102, JP16K17975, and JP17H03137. We would like to thank Editage ( www.editage.jp) for English language editing.

References

1. 

G. N. Lewis, D. Lipkin and T. T. Magel, “Reversible photochemical processes in rigid media. A study of the phosphorescent state,” J. Am. Chem. Soc., 63 (11), 3005 –3018 (1941). http://dx.doi.org/10.1021/ja01856a043 JACSAT 0002-7863 Google Scholar

2. 

C. A. Parker and C. G. Hatchard, “Triplet-singlet emission in fluid solutions—phosphorescence of eosin,” Trans. Faraday Soc., 57 (11), 1894 –1904 (1961). http://dx.doi.org/10.1039/tf9615701894 TFSOA4 0014-7672 Google Scholar

3. 

H. Uoyama et al., “Highly efficient organic light-emitting diodes from delayed fluorescence,” Nature, 492 (7428), 234 –238 (2012). http://dx.doi.org/10.1038/nature11687 Google Scholar

4. 

C. Adachi, “Third-generation organic electroluminescence materials,” Jpn. J. Appl. Phys., 53 (6), 060101 (2014). http://dx.doi.org/10.7567/JJAP.53.060101 Google Scholar

5. 

D. Volz, “Review of organic light-emitting diodes with thermally activated delayed fluorescence emitters for energy-efficient sustainable light sources and displays,” J. Photonics Energy, 6 020901 (2016). http://dx.doi.org/10.1117/1.JPE.6.020901 Google Scholar

6. 

R. Komatsu et al., “Light-blue thermally activated delayed fluorescent emitters realizing a high external quantum efficiency of 25% and unprecedented low drive voltages in OLEDs,” J. Mater. Chem. C, 4 (12), 2274 –2278 (2016). http://dx.doi.org/10.1039/C5TC04057D Google Scholar

7. 

Y. J. Xie and Z. Li, “Thermally activated delayed fluorescent polymers,” J. Polym. Sci. Polym. Chem., 55 (4), 575 –584 (2017). http://dx.doi.org/10.1002/pola.28448 Google Scholar

8. 

Z. Yang et al., “Recent advances in organic thermally activated delayed fluorescence materials,” Chem. Soc. Rev., 46 (3), 915 –1016 (2017). http://dx.doi.org/10.1039/C6CS00368K CSRVBR 0306-0012 Google Scholar

9. 

M. Y. Wong and E. Zysman-Colman, “Purely organic thermally activated delayed fluorescence materials for organic light-emitting diodes,” Adv. Mater., 29 1605444 (2017). http://dx.doi.org/10.1002/adma.201605444 ADVMEW 0935-9648 Google Scholar

10. 

T. Miwa et al., “Blue organic light-emitting diodes realizing external quantum efficiency over 25% using thermally activated delayed fluorescence emitters,” Sci. Rep., 7 (1), 284 (2017). http://dx.doi.org/10.1038/s41598-017-00368-5 SRCEC3 2045-2322 Google Scholar

11. 

H. Nakanotani et al., “Promising operational stability of high-efficiency organic light-emitting diodes based on thermally activated delayed fluorescence,” Sci. Rep., 3 2127 (2013). http://dx.doi.org/10.1038/srep02127 SRCEC3 2045-2322 Google Scholar

12. 

H. Kaji et al., “Purely organic electroluminescent material realizing 100% conversion from electricity to light,” Nat. Commun., 6 8476 (2015). http://dx.doi.org/10.1038/ncomms9476 NCAOBW 2041-1723 Google Scholar

13. 

T. Furukawa et al., “Dual enhancement of electroluminescence efficiency and operational stability by rapid upconversion of triplet excitons in OLEDs,” Sci. Rep., 5 8429 (2015). http://dx.doi.org/10.1038/srep08429 SRCEC3 2045-2322 Google Scholar

14. 

D. P.-K. Tsang, T. Matsushima and C. Adachi, “Operational stability enhancement in organic light-emitting diodes with ultrathin Liq interlayers,” Sci. Rep., 6 22463 (2016). http://dx.doi.org/10.1038/srep22463 SRCEC3 2045-2322 Google Scholar

15. 

F. B. Dias et al., “Triplet harvesting with 100% efficiency by way of thermally activated delayed fluorescence in charge transfer OLED emitters,” Adv. Mater., 25 (27), 3707 –3714 (2013). http://dx.doi.org/10.1002/adma.v25.27 ADVMEW 0935-9648 Google Scholar

16. 

X. K. Chen et al., “Nature of highly efficient thermally activated delayed fluorescence in organic light-emitting diode emitters: nonadiabatic effect between excited states,” J. Phys. Chem. C, 119 (18), 9728 –9733 (2015). http://dx.doi.org/10.1021/acs.jpcc.5b00276 JPCCCK 1932-7447 Google Scholar

17. 

T. Ogiwara, Y. Wakikawa and T. Ikoma, “Mechanism of intersystem crossing of thermally activated delayed fluorescence molecules,” J. Phys. Chem. A, 119 (14), 3415 –3418 (2015). http://dx.doi.org/10.1021/acs.jpca.5b02253 JPCAFH 1089-5639 Google Scholar

18. 

J. Gibson, A. P. Monkman and T. J. Penfold, “The importance of vibronic coupling for efficient reverse intersystem crossing in thermally activated delayed fluorescence molecules,” Chem. Phys. Chem., 17 (19), 2956 –2961 (2016). http://dx.doi.org/10.1002/cphc.201600662 CPCHFT 1439-4235 Google Scholar

19. 

C. M. Marian, “Mechanism of the triplet-to-singlet upconversion in the assistant dopant ACRXTN,” J. Phys. Chem. C, 120 (7), 3715 –3721 (2016). http://dx.doi.org/10.1021/acs.jpcc.6b00060 JPCCCK 1932-7447 Google Scholar

20. 

F. B. Dias et al., “The role of local triplet excited states and D-A relative orientation in thermally activated delayed fluorescence: photophysics and devices,” Adv. Sci., 3 1600080 (2016). http://dx.doi.org/10.1002/advs.v3.12 Google Scholar

21. 

P. K. Samanta et al., “Up-conversion intersystem crossing rates in organic emitters for thermally activated delayed fluorescence: impact of the nature of singlet vs triplet excited states,” J. Am. Chem. Soc., 139 (11), 4042 –4051 (2017). http://dx.doi.org/10.1021/jacs.6b12124 JACSAT 0002-7863 Google Scholar

22. 

T. Kobayashi et al., “Contributions of a higher triplet excited state to the emission properties of a thermally activated delayed-fluorescence emitter,” Phys. Rev. Appl., 7 (3), 034002 (2017). http://dx.doi.org/10.1103/PhysRevApplied.7.034002 PRAHB2 2331-7019 Google Scholar

23. 

M. K. Etherington et al., “Regio- and conformational isomerization critical to design of efficient thermally-activated delayed fluorescence emitters,” Nat. Commun., 8 14987 (2017). http://dx.doi.org/10.1038/ncomms14987 NCAOBW 2041-1723 Google Scholar

24. 

T. Hosokai et al., “Evidence and mechanism of efficient thermally activated delayed fluorescence promoted by delocalized excited states,” Sci. Adv., 3 (5), e1603282 (2017). http://dx.doi.org/10.1126/sciadv.1603282 STAMCV 1468-6996 Google Scholar

25. 

Q. Peng et al., “Theoretical study of conversion and decay processes of excited triplet and singlet states in a thermally activated delayed fluorescence molecule,” J. Phys. Chem. C, 121 (25), 13448 –13456 (2017). http://dx.doi.org/10.1021/acs.jpcc.7b00692 JPCCCK 1932-7447 Google Scholar

26. 

T. Sato et al., “Fluorescence via reverse intersystem crossing from higher triplet states in a bisanthracene derivative,” Sci. Rep., 7 (1), 4820 (2017). http://dx.doi.org/10.1038/s41598-017-05007-7 SRCEC3 2045-2322 Google Scholar

27. 

K. Masui, H. Nakanotani and C. Adachi, “Analysis of exciton annihilation in high-efficiency sky-blue organic light-emitting diodes with thermally activated delayed fluorescence,” Org. Electron., 14 (11), 2721 –2726 (2013). http://dx.doi.org/10.1016/j.orgel.2013.07.010 OERLAU 1566-1199 Google Scholar

28. 

M. Inoue et al., “Effect of reverse intersystem crossing rate to suppress efficiency roll-off in organic light-emitting diodes with thermally activated delayed fluorescence emitters,” Chem. Phys. Lett., 644 62 –67 (2016). http://dx.doi.org/10.1016/j.cplett.2015.11.042 CHPLBC 0009-2614 Google Scholar

29. 

A. S. D. Sandanayaka et al., “Toward continuous-wave operation of organic semiconductor lasers,” Sci. Adv., 3 (4), e1602570 (2017). http://dx.doi.org/10.1126/sciadv.1602570 STAMCV 1468-6996 Google Scholar

30. 

J. R. Lakowicz, Principles of Fluorescence Spectroscopy, 3rd ed.Springer, New York (2006). Google Scholar

31. 

R. Katoh et al., “Ultrafast relaxation as a possible limiting factor of electron injection efficiency in black dye sensitized nanocrystalline TiO2 films,” J. Phys. Chem. C, 116 (42), 22301 –22306 (2012). http://dx.doi.org/10.1021/jp302768q JPCCCK 1932-7447 Google Scholar

32. 

T. Yoshihara, S. I. Druzhinin and K. A. Zachariasse, “Fast intramolecular charge transfer with a planar rigidized electron donor/acceptor molecule,” J. Am. Chem. Soc., 126 (27), 8535 –8539 (2004). http://dx.doi.org/10.1021/ja049809s JACSAT 0002-7863 Google Scholar

33. 

Q. Zhang et al., “Design of efficient thermally activated delayed fluorescence materials for pure blue organic light emitting diodes,” J. Am. Chem. Soc., 134 (36), 14706 –14709 (2012). http://dx.doi.org/10.1021/ja306538w JACSAT 0002-7863 Google Scholar

34. 

R. Ishimatsu et al., “Solvent effect on thermally activated delayed fluorescence by 1,2,3,5-tetrakis(carbazol-9-yl)-4, 6-dicyanobenzene,” J. Phys. Chem. A, 117 (27), 5607 –5612 (2013). http://dx.doi.org/10.1021/jp404120s JPCAFH 1089-5639 Google Scholar

35. 

H. Hayashi and S. Nagakura, “The E.S.R. and phosphorescence spectra of some dicyanobenzene complexes with methyl-substituted benzenes,” Mol. Phys., 19 (1), 45 –53 (1970). http://dx.doi.org/10.1080/00268977000100991 MOPHAM 0026-8976 Google Scholar

Biography

Takuya Hosokai is a scientist at the National Institute of Advanced Industrial Science and Technology, in Japan. He received his BS and MS degrees in engineering from Niigata University and Yokohama National University in 2003 and 2005, respectively, and his PhD in engineering from Chiba University in 2008. He was a postdoc at Universität Tübingen from 2008 to 2010, where he received a research scholarship from the Alexander von Humboldt Foundation. He is the author of more than 50 journal publications and has written three book chapters. His current research interests include optical and electronic properties of organic semiconducting molecules, and the development of new experimental instruments.

Biographies for the other authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Takuya Hosokai, Hiroki Noda, Hajime Nakanotani, Takanori Nawata, Yasuo Nakayama, Hiroyuki Matsuzaki, and Chihaya Adachi "Solvent-dependent investigation of carbazole benzonitrile derivatives: does the 3LE−1CT energy gap facilitate thermally activated delayed fluorescence?," Journal of Photonics for Energy 8(3), 032102 (2 February 2018). https://doi.org/10.1117/1.JPE.8.032102
Received: 16 November 2017; Accepted: 10 January 2018; Published: 2 February 2018
Lens.org Logo
CITATIONS
Cited by 27 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Luminescence

Molecules

Phosphorescence

Absorption

Absorbance

Electronics

Organic light emitting diodes

Back to Top