Ikki
Kamiyoshi
a,
Yuki
Kojima
a,
Shibo
Xu
b,
Kosuke
Yasui
ab,
Yuji
Nishii
ab and
Koji
Hirano
*ab
aDepartment of Applied Chemistry, Graduate School of Engineering, Osaka University, Suita, Osaka 565-0871, Japan. E-mail: k_hirano@chem.eng.osaka-u.ac.jp
bInnovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, Suita, Osaka 565-0871, Japan
First published on 14th November 2024
A photo-promoted oxidative cyclization, that is, the Mallory reaction of 2,3-diarylbenzophopholes has been developed. With the assistance of Bi(OTf)3 Lewis acid, the reaction proceeds smoothly under visible light irradiation even without any external oxidants. The newly developed dehydrogenative conditions are compatible with various functional groups and substitution patterns, which enables the streamlined synthesis of highly condensed dibenzophosphole derivatives of potent interest in material chemistry. Moreover, experimental and computational studies unveil the detailed reaction mechanism. The preliminary optoelectronic properties of some newly synthesized compounds are also demonstrated.
On the other hand, in recent years our research group serendipitously found that 3-indolylphosphole underwent dehydrogenative cyclization even under visible light irradiation.6 This phenomenon was specific to a phosphole bearing the strongly electron-donating indole substituent at the C3 position, suggesting that the substrate scope can be extended by suitable electronic perturbation with the assistance of external additives. Herein, we report the significant substrate extension: the Bi(OTf)3-assisted Mallory reaction of 2,3-diarylbenzophospholes under blue LED irradiation (456 nm) has been developed (Scheme 1c). The reaction conditions accommodate electron-donating and electron-withdrawing aryl groups as well as heteroaryl groups on the phosphole core. In addition, mild reaction conditions with visible light of lower energy are tolerant of a wide range of functional groups such as the Ar–Br moiety, which can be further transformed. Moreover, even in the absence of any external oxidants, the reaction proceeds smoothly with concomitant evolution of molecular hydrogen. Experimental and computational mechanistic studies and preliminary optoelectronic properties of some newly synthesized condensed phosphole derivatives are also described.
Entry | Additives | Yield of 2ab (%) |
---|---|---|
a Conditions: 1a (0.050 mmol), additives (0.050 mmol), MeCN (1.0 mL), blue LED (456 nm, 40 W), ambient temperature (40–50 °C under light irradiation), 22 h, N2. b Estimated by using 31P{1H} NMR based on 0.050 mmol with P(O)(OEt)3 as the internal standard. Isolated yield is in parentheses. c In the dark at 50 °C. Tf = trifluoromethanesulfonyl, TFA = trifluoroacetic acid, PTSA = p-toluenesulfonic acid, and Ac = acetyl. | ||
1 | Bi(OTf)3 | 33 |
2 | Al(OTf)3 | 49 |
3 | In(OTf)3 | 68 |
4 | Sc(OTf)3 | 34 |
5 | TFA | 40 |
6 | (PhO)2P(O)(OH) | 20 |
7 | TfOH | 0 |
8 | PTSA | 0 |
9 | HCl | 0 |
10 | AcOH | 0 |
11 | Bi(OTf)3, NaHCO3 | (>99) |
12 | Bi(OTf)3, Na2CO3 | <5 |
13 | Al(OTf)3, NaHCO3 | 38 |
14 | In(OTf)3, NaHCO3 | 31 |
15 | Sc(OTf)3, NaHCO3 | 8 |
16c | Bi(OTf)3, NaHCO3 | 0 |
17 | None | 0 |
Combined with our previously developed convergent synthetic methods of 2,3-disubstituted benzophospholes, the visible-light-promoted Mallory reaction can offer a modular approach to the highly condensed dibenzophospholes with structural diversity. One is the radical annulation–Mallory reaction sequence (Scheme 2a): a variety of 2,3-diarylbenzophospholes 1a–h readily prepared from diphenylphosphine oxide and symmetrical diaryl alkynes7 underwent the Mallory reaction smoothly (Scheme 2a). Electron-neutral (2b and 2c), -rich (2d), and -deficient (2e) substrates were all tolerated. Owing to the visible light of relatively lower energy, the reaction conditions were compatible with the Ar–Br bond (2f), which is a versatile handle in downstream transformation. Actually, the Pd-catalysed double Buchwald–Hartwig amination of 2f with carbazole and phenoxazine8 enabled concise synthesis of the dibenzophosphole-based donor–acceptor systems 3fa and 3fb in 83% and 27% yield, respectively (Scheme 3). In the case of 2-naphthyl-substituted benzophosphole 1g, dehydrogenative cyclization occurred regioselectively to deliver the helical 2g in an almost quantitative yield. Its structure was unambiguously confirmed by X-ray analysis (CCDC 2377855).† The heteroaromatic benzothiophene substituent could also be employed, and the corresponding P,S-doped triphenylene 2h was obtained in an acceptable yield. Combination with the phosphorus-cation-mediated annulation of the t-Bu-substituted diphenyl phosphine oxide9a successfully provided dibenzophosphole 2i with the t-Bu-C6H4 substituent on the phosphorus, indicating that the diverse substitution pattern is accessible. The reaction also proceeded on a 1.0 mmol scale with synthetically useful efficacy (2a).
![]() | ||
Scheme 3 Double Buchwald–Hartwig amination of dibrominated benzophosphole 2f with carbazole and phenoxazine. See the ESI† for detailed reaction conditions. |
The phosphorus-dication-mediated annulation9b and successive Pd-catalysed C–H arylation10 provided facile access to the benzophospholes 1j–m with two different aromatic rings at the C2 and C3 positions, which were also amenable substrates in the Mallory reaction (Scheme 2b). The highly electron-donating trimethoxyphenyl (2j) and methylenedioxyphenyl (2k) groups at the C2 position were well tolerated under specific reaction conditions. Notably, exclusive regioselectivity was observed in the latter case. On the other hand, the dibenzothiophene-substituted substrate 1l afforded a regioisomeric mixture of helical 2l and linear 2l′ (CCDC 2377856).† The Lewis basic quinoline-fused dibenzophosphole 2m was also obtained albeit with a moderate yield.
The Mallory reaction was also applicable to the C2-alkenyl-C3-arylbenzophospholes 1n and 1o, which were also easily synthesized via Pd-catalysed C–H alkenylation11 (Scheme 2c). In the case of 1n, we successfully obtained the 5,7-diphenyldibenzo[b,e]phosphindole 7-oxide structure 2n, which is employed as the OLED device for improvement of emission efficiency.12 With the MeO-substituted substrate 1o, the partial demethylation of the MeO group competitively occurred, but a good combined yield of 2o and 2o-H was observed.
To gain insight into the reaction mechanism, several experiments were performed next. First, we investigated potential oxidants in the present Mallory reaction (Scheme 4a). Zhao recently reported that residual oxygen in solvent is the actual external oxidant in the related UV-light-promoted Mallory reaction, which was supported by no conversion under rigorously deaerated conditions using freeze–pump–thaw cycling.5 Thus, our Bi(OTf)3-assisted, visible-light-promoted Mallory reaction was also conducted by using the same deaeration technique. Surprisingly, the reaction proceeded smoothly without any difficulty, and 2a was formed in 96% yield comparable with that under the standard nitrogen conditions (>99% yield, Table 1, entry 11). In addition, a lower yield was observed in air. These results suggest that oxygen in air unlikely works as the external oxidant. Finally, we successfully observed evolution of molecular hydrogen by GC after the reaction, thus indicating that the present Mallory reaction is acceptor-less, and a truly dehydrogenative reaction. To check whether the reaction is operated in the singlet or triplet excited state, we then carried out the reaction of 1a in the presence of 2,5-dimethyl-2,4-hexadiene, which is known as a triplet quencher,13 but 2a was still formed in a good yield (Scheme 4b). Even with (E)-methyl cinnamate, which undergoes E to Z isomerization by triplet–triplet energy transfer,141a was converted to 2a efficiently, and cinnamate was recovered with the maintenance of its (E)-configuration. Although the possibility of a triplet state intermediacy cannot be completely excluded, given the effective quenching even in the case of intramolecular cyclization,15 the aforementioned outcomes support the reaction progress via the singlet excited state rather than the triplet excited state. Information about the rate-determining step was obtained from the deuterium-labeling experiment with 1a-d10 (Scheme 4c): the kinetic isotope effect (KIE) value from the parallel reaction was 2.90, thus suggesting the rate-limiting C–H cleavage. Finally, the role of Bi(OTf)3 was examined by using NMR and UV-vis absorption spectra (Scheme 4). Upon mixing of 1a and Bi(OTf)3 in CDCl3, some 1H and 31P{1H} NMR signals were shifted to the lower field. The absorption spectrum was also red-shifted, upon exposure of 1a to Bi(OTf)3, to the visible light region. The colour change in the solution could also be confirmed visually. Apparently, the complexation between the phosphine oxide moiety of 1a and Bi(OTf)3 enables excitation even with visible light.
![]() | ||
Scheme 5 Computational mechanistic studies. Summary of the level of theory: M06-2X/6-311+G(d,p)&SDD/PCM(MeCN)//M06-2X/6-31G(d)&LanL2DZ. |
Additional computational mechanistic studies were performed by using density functional theory (DFT) calculations (Scheme 5). All calculations were carried out using the Gaussian 16 program.16 The singlet ground-state (S0) and singlet exited-state (S1) geometries were optimized by the DFT and time-dependent density functional theory (TD-DFT) methods with the M06-2X functional and a standard 6-31G(d) basis set (LanL2DZ basis set for Bi). The M06-2X functional is a high-nonlocality functional with double the amount of nonlocal exchange (2X), with reliable performance for thermochemistry, hydrogen bonding, kinetics, and weak interactions.17 The optimized molecular structures were verified by vibrational analysis; equilibrium structures did not have imaginary frequencies and transition state structures had only one imaginary frequency. The intrinsic reaction coordinate (IRC) calculations were carried out to check the transition state leading to the reactant and the product. Single-point energies were calculated using the 6-311+G(d,p) basis set (SDD basis set for Bi), and the solvent effect of MeCN was taken into account by using the integral equation formalism PCM (IEF-PCM). We first optimized the molecular structures of 1a and Bi(OTf)3-coordinated 1a in the S0 state and calculated their HOMO and LUMO levels (Scheme 5a). Upon the coordination of Bi(OTf)3, the smaller HOMO–LUMO gap is obtained particularly by lowering the LUMO level, which is in good agreement with the NMR and UV-vis absorption spectral changes observed in Scheme 4d.18 In addition, the Bi(OTf)3 coordination is an energetically favoured, exothermic process (ΔG = −28.0 kcal mol−1). Thus, Bi(OTf)3-coordinated 1a (intA_s0) is a rational starting point in the present Mallory reaction (Scheme 5b). Consistent with no conversion under dark conditions (Table 1, entry 16), the activation barrier for direct electrocyclization in the ground state (tsAB_s0) is +72.9 kcal mol−1, which can exclude the thermal reaction pathway. On the other hand, the S1 intermediate intA_s1 undergoes conrotatory electrocyclization to form intB_s0viaintB_s1 with a reasonable energy change (+38.0 to +59.4 to +15.4 kcal mol−1). The direct elimination of vic-hydrogens in intB_s0 is difficult because they are anti to each other. Thus, enolization-induced epimerization occurs through intC_s0 to afford the corresponding syn-isomer intD_s0. The process is uphill, but subsequent spontaneous elimination of molecular hydrogen into the gas phase and rearomatization into intE are strong driving forces for the reaction process. The overall energy profile reveals that the rate-limiting step is the C–H cleaved enolization from intB_s0 to intC_s0, which is also in accordance with the primary KIE value observed in Scheme 4c.
The optical properties of some newly obtained compounds in Schemes 2 and 3 (2j, 2k, 2l, 2l′, 3fa, and 3fb) were preliminary surveyed in a solution state (1.0 × 10−5 M CHCl3 solution). The data of absorption/emission properties (λabs/λem) and fluorescence quantum yields (ΦF), also involving those of 1a and 2a as the references, are summarized in Fig. 1 and Table 2. Compared to the uncyclized 1a, the condensed 2a–2l′ all showed bathochromic shifts of their λabs values (379–412 nm) probably because of effective π-extension. On the other hand, impacts on the emission maxima λem were little or negligible. Relatively high quantum yields were observed in the push–pull-type alkoxy-substituted 2j and 2k, while higher fused 2l and 2l′ containing the dibenzothiophene core showed lower emission efficiency. In general, the smaller Stokes shifts were obtained from all compounds owing to their rigid planar structures. The carbazole-incorporated dibenzophosphole 3fa exhibited both absorption and emission maxima in the longer wavelength regions, which is reflected by the strongly electron-donating nature of the carbazole unit. Additionally, a very high quantum yield was obtained. The introduction of phenoxazine induced much larger red shifts of λabs and λem albeit with poor emission efficiency (3fb).
Compd | λ abs (nm) | λ em (nm) | Φ F | Δνd/cm−1 |
---|---|---|---|---|
a Measured in 1.0 × 10−5 M solution of CHCl3. b Excited at 1a (249 nm), 2a (254 nm), 2j (260 nm), 2k (260 nm), 2l (270 nm), 2l′ (262 nm), 3fa (244 nm), and 3fb (298 nm). c Absolute fluorescence quantum yields. d Stokes shifts. e The optical data of 2a were taken from ref. 10. | ||||
1a | 249, 342 | 438 | 0.11 | 6409 |
2a | 254, 315, 364, 382 | 425 | 0.56 | 2649 |
2j | 260, 379 | 448 | 0.58 | 4064 |
2k | 260, 403 | 434 | 0.75 | 1772 |
2l | 212, 270, 324, 392 | 443 | 0.16 | 2937 |
2l′ | 262, 310, 344, 391, 412 | 432 | 0.16 | 1124 |
3fa | 244, 291, 341, 396 | 473 | >0.99 | 4111 |
3fb | 240, 258, 298, 317, 440 | 620 | 0.04 | 6598 |
We also examined the electrochemical properties by cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in MeCN with tetrabutylammonium hexafluorophosphate (Bu4NPF6) as an electrolyte versus ferrocene/ferrocenium ions (Fc/Fc+) (Fig. S14–S19†), and their HOMO and LUMO levels were estimated according to the first oxidation potentials and the optical band gaps (Eoptg). The data are summarized in Table 3. The CV of all compounds showed irreversible oxidation waves, and thus Eox values were determined by DPV. The electron-donating alkoxy (2j and 2k) and amino (3fa and 3fb) substituents largely shifted Eox values in a negative direction from those of the parent 2a. Notably, compared to the carbazole-substituted 3fa, the phenoxazine-substituted 3fb exhibited a higher-lying HOMO level but an almost identical LUMO level, thus suggesting its larger intramolecular charge transfer ability. On the other hand, the structural isomers 2l and 2l′ showed almost the same values for all parameters: their electrochemical properties are less dependent on the orientation of fused dibenzothiophene.
Compd | λabsonseta (nm) | E optg (eV) | E ox (V) | E HOMO (eV) | E LUMO (eV) |
---|---|---|---|---|---|
a Measured in CHCl3. b Determined from the onset of the absorption spectra. c Performed in MeCN in the presence of Bu4NPF6. v = 0.10 V s−1. Values determined by DPV, versus Fc/Fc+. d The approximation for the Fc/Fc+ level is −4.8 eV versus vacuum: EHOMO = −4.8 − Eox. e Estimated from EHOMO and Eoptg: ELUMO = EHOMO + Eoptg. f The data of 2a were taken from ref. 10. | |||||
2a | 397 | 3.12 | 1.32 | −6.12 | −3.00 |
2j | 422 | 2.94 | 0.91 | −5.71 | −2.77 |
2k | 424 | 2.92 | 0.99 | −5.79 | −2.87 |
2l | 429 | 2.89 | 1.08 | −5.88 | −2.99 |
2l′ | 428 | 2.90 | 1.07 | −5.87 | −2.97 |
3fa | 445 | 2.79 | 0.72 | −5.52 | −2.73 |
3fb | 521 | 2.38 | 0.32 | −5.12 | −2.74 |
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2377855 and 2377856. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4sc05657d |
This journal is © The Royal Society of Chemistry 2024 |