Open Access Article
Bo
Yang
a,
Suqiong
Yan
a,
Chengbo
Li
c,
Hui
Ma
a,
Fanda
Feng
a,
Yuan
Zhang
a and
Wei
Huang
*ab
aState Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, P. R. China. E-mail: whuang@nju.edu.cn
bShenzhen Research Institute of Nanjing University, Shenzhen 51805, P. R. China
cSchool of Materials and Energy, University of Electronic Science and Technology of China, Chengdu 610000, P. R. China
First published on 24th August 2023
Transition metal mediated C–X (X = H, halogen) bond activation provides an impressive protocol for building polyaromatic hydrocarbons (PAHs) in C–C bond coupling and annulation; however, mimicking both the reaction model and Lewis acid mediator simultaneously in a hetero-PAH system for selective C–P bond cleavage faces unsolved challenges. At present, developing the C–P bond activation protocol of the phosphonic backbone using noble-metal complexes is a predominant passway for the construction of phosphine catalysts and P-center redox-dependent photoelectric semiconductors, but non-noble metal triggered methods are still elusive. Herein, we report Mn(III)-mediated C–P bond activation and intramolecular cyclization of diphosphines by a redox-directed radical phosphonium process, generating phosphahelicene cations or phosphoniums with nice regioselectivity and substrate universality under mild conditions. Experiments and theoretical calculations revealed the existence of the unusual radical mechanism and electron-deficient character of novel phosphahelicenes. These rigid quaternary bonding skeletons facilitated versatile fluorescence with good tunability and excellent efficiency. Moreover, the enantiomerically enriched crystals of phosphahelicenes emitted intense circularly polarized luminescence (CPL). Notably, the modulated CPL of racemic phosphahelicenes was induced by chiral transmission in the cholesteric mesophase, showing ultrahigh asymmetry factors of CPL (+0.51, −0.48). Our findings provide a new approach for the design of emissive phosphahelicenes towards chiral emitters and synthesized precursors.
Morandi's group pioneered the C–P bond activation of diphosphines via Pd2(dba)3-catalyzed metathesis (Scheme 1a), and the PPh3 fragment was eliminated for the cyclopalladated intermediate. Finally, novel neutral phospha[5]helicenes (PIII) were generated and further racemic phospha[5]helicene (PV) oxides were isolated.16 Subsequently, Wang's group modified the reaction to break the intramolecular cyclization of pre-activated phosphoniums.15 In addition, Cu(OTf)2-mediated C–H bond activation at 8,8′-positions in binaphthyl produced an achiral bisphosphonium with a low yield of 24%.20 This method spread in the laboratory for the preparation of achiral phosphoniums.21,22 Theoretically, another potential cationic phosphahelicene might be produced after eliminating the PPh2 unit and rebuilding the C–P bond at the 2′-position. However, this valued phosphahelicene cation ([1b]+) and its properties have not been captured and explored (Scheme 3). These above described electron-deficient cationic phosphahelicenes that integrated an active C–P bond and a prochiral axis to enhance the oxidative addition efficacy endowed the enantioselective cross-coupling from C–P to C–C bonds with excellent stereoselectivity and regioselectivity.23 We envisioned that the tunability of metal sources in the pursuit of altering the activation of radical cations and elimination pathways of the intermediates would change the reactivity and selectivity of cyclization.12,16
Recently, heterohelicenes have shown arresting renaissance due to their tunable electronic structures, making them widely used in asymmetric catalysis and optoelectronics.24–28 Notably, the electronic factors are highly dependent on the oxidation state of the phosphorus center in phosphahelicenes. Most neutral phosphahelicenes (PIII) yielded lower quantum efficiency and insufficient CPL because of high nonradiative rates and the fluxional molecular conformation in the excited states (i.e., analogues phospha[5]helicenes 1 and 2 are not luminescent, Scheme 1c).29–33 To overcome this drawback, we speculated that introducing a quaternary bonding environment for the P-atom could increase the steric hindrance to reset the electronic transition nature in phosphahelicenes, thereby inhibiting the nonradiative decay of emitters in excited states.22,34 Furthermore, the cationic phosphahelicene has high stability and electron-defect character in the quaternary C–P bonding environment, providing a new opportunity for the design of chiral donor–acceptor (D–A) modules.35–37 On the other hand, exploring chiral transmission toward increased CPL of racemic [5]helicenes is still rare.38,39 Hence, developing highly emissive cationic phosphahelicene scaffolds and chiral self-assembly tactics is an essential extension of heterohelicenes for photoelectric devices and synthesized precursors.
Unprecedentedly, we reported a novel MnX3 (X = Cl, Br) mediated protocol to construct cationic phosphahelicenes and phosphoniums from diphosphines (Scheme 1b) via intramolecular cyclization with good regioselectivity, substrate scope, and yields (up to 92%). Mechanistically, redox-directed C–P bond cleavage and annulation were confirmed by a radical phosphonium path of the reaction. The intense fluorescence of the products can be regulated from blue to orange (ΦPLQY up to 77%). Importantly, modulated CPL was obtained for this series of emissive phosphahelicenes in a ternary cholesteric mesophase by chiral transmission, showing ultrahigh asymmetry factors (+0.51, −0.48). Mechanistic study demonstrated that the increased CPL critically depended on chiral fluorescence resonance energy transfer (FRET) and helical superstructures rather than selective Bragg reflection.
| Entry | Mediator | Solvent | T (°C) | Yieldc (%) |
|---|---|---|---|---|
a Reaction conditions: BINAP (0.16 mmol), metal-mediator (3.0 eq.). aMixed-solvent (Vchloroform : VEtOH = 2 : 1, 10 mL), under an O2 atmosphere (1 bar) at corresponding temperature for 5–10 minutes.
b 10 mol%.
c Isolated yields for helicenes and BINAPOs (in parentheses).
d ND represents not detected.
e Ar instead of O2.
|
||||
| 1 | MnCl2 | CHCl3 : EtOH |
30 | 56(37), NDe |
| 2 | MnCl2 | CHCl3 : EtOH |
60 | 51(40) |
| 3 | MnCl2 | CHCl3 : EtOH |
80 | 52(40) |
| 4 | MnBr2 | CHCl3 : EtOH |
30 | 54(42) |
| 5 | MnI2 | CHCl3 : EtOH |
30 | Trace |
| 6 | MnCO3 | CHCl3 : EtOH |
80 | NDd |
| 7 | Mn(OTf)2 | CHCl3 : EtOH |
80 | NDd |
| 8 | Mn(acac)3 | CHCl3 : EtOH |
80 | NDd |
| 9 | Mn(OAc)3 | CHCl3 : EtOH |
80 | NDd |
| 10 | FeCl3 | CHCl3 : EtOH |
80 | NDd |
| 11 | CoCl2 | CHCl3 : EtOH |
80 | NDd |
| 12 | NiCl2 | CHCl3 : EtOH |
80 | NDd |
| 13 | CuCl2 | CHCl3 : EtOH |
80 | NDd |
| 14 | MnCl2 | Toluene | 80 | Trace |
| 15 | MnCl2 | Chloroform | 80 | 38(35) |
| 16 | MnCl2 | Ethanol | 80 | <5(11) |
| 17 | MnCl2 | THF | 80 | 42(47) |
| 18 | MnCl2 | Acetonitrile | 80 | 29(34) |
| 19 | MnCl2b | CHCl3 : EtOH |
30 | Trace |
Based on the optimized conditions in hand, the substrate scope was investigated for BINAPs/BIPHEPs (Scheme 2). Derivatives BINAP-4-Me and BINAP-3,5-Me2 with methyl groups at different positions in phenyls converted into [2b]+[Cl]− and [3b]+[Cl]− with 54% and 56% yields. Partial BINAPs showed slow cyclization at room temperature due to lower activity, which required proper heating (50–80 °C, Scheme 2). BINAPs with electron-donating methyl groups at ortho, meta, and para-positions converted into [11b]+[Cl]−, [12b]+[Cl]−, and [13b]+[Cl]− with depressed yields (37–42%), but the BINAPOs could be reduced and transformed into initial BINAPs by HSiCl3 and utilized next time. Besides, ethyl, methoxyl, tertiary butyl, and mesitylene substituted BINAPs displayed higher yields (61–75%). Encouragingly, electron-deficient BINAPs (fluorine, trifluoromethyl, and ester) showed the highest yields (up to 85%). It was worth noting that the extensional phenyls, biphenyls, and naphthyls had an inconspicuous influence on intramolecular cyclization. Similarly, biphenyl diphosphines (BIPHEPs) cyclized into corresponding phosphoniums with moderate yields (39–67%). Unfortunately, the P,O-embedded phosphonium [8b]+[Cl]− did not generate in this reaction, which was related to the flexible aryl ether motion of the DPEPhos substrate and longer P–P distance (4.88 Å) than that of BINAP (4.20 Å), leading to a different chelation environment for the intermediate (Fig. S7†). Furthermore, 5,5′-modified BINAPs showed satisfactory tolerance (40–73%) equal to those of 6,6′-modified BINAPs. Nevertheless, the 7,7′-modified BINAPs exhibited excellent cyclization, getting the highest yields (68–92%) of phospha[5]helicenes, which could be relevant to variational electronic structures and an elevated competitive rate between cyclization and oxidation. Finally, gram-scale reactions were performed with good yields (Scheme 2). To the best of our knowledge, this method represented the first C–P bond activation discovery in diphosphines using non-noble metal mediators. This family of helicenium products could be transformed into asymmetric monophosphines by Pd-catalyzed stereoselective cleavage of the C–P bond and cross-coupling according to a recent report.23
Indeed, this cyclization was highly stereospecific, where the product chirality inherited the BINAPs. However, the enantiomeric purity decreased in solution due to a low interconversion barrier at room temperature.23,42 Enantiomerically enriched [2b]+[Cl]− and [2b]+[BF4]− were acquired via rapid synthesis, chromatographic purification, and their single-crystal growth at low temperature to guarantee adequate enantiomeric purity. Combined with the observations of SCXRD and uniform morphologies (Fig. S8, S9, S16, S17, and S21†), these proofs affirmed the formation of homochiral single crystals. The homochiral crystals were used for the kinetic study of chirality. Initially, we attempted to evaluate the ee values of compounds using chiral HPLC (at 25 °C) but failed. To this end, we then employed the temperature and time-dependence of the rate constant to calculate the ee values and thermodynamic parameters of the racemization process via CD (circular dichroism) spectra (see the ESI for detailed determination, Fig. S1†). According to the CD ellipticities between homochiral crystals and pristine products in solution, the average ee value (for 5 parallel samples) of pristine product [2b]+[Cl]− was about 73%. Finally, as a summary of Eyring plotting and calculations for [2b]+[Cl]−, ΔH‡ = 12.5 kcal mol−1 and ΔS‡ = −25.4 cal mol−1 K−1 were determined, and ΔG‡exp = 20.1 kcal mol−1, k = 1.5 × 10−4 s−1, and t1/2 (half-life) = 77.1 min at 298.15 K were successfully elucidated, whose ΔG‡exp value was smaller than that of carbon[5]helicene (∼25 kcal mol−1 at 298.15 K).43–45 The ΔG‡exp value was slightly below the margin of the activation barrier (∼22.2 kcal mol−1) required for conformational atropo-enantiomers to be resolvable.46 In addition, the interconversion of helical conformation was also considered via density functional theory (DFT) calculations, and the isomerization crossed over the saddle-shaped transition state with a low isomerization barrier (ΔG‡cal = 24.1 kcal mol−1, at the B3LYP/6-31G* level) of [(P)/(M)-2b]+ (Fig. S38†). The degraded ee value observed in our experiments was reasonable, and the experimental ΔG‡cal value was smaller by 4.0 kcal mol−1 than the theoretical one. The reason for the latter may be the problem of the accuracy of the basis set/functionality.47 The lower ΔG‡cal 18.6 kcal mol−1 of [(P)/(M)-1b]+ facilitated the loss of stereochemical purity completely and the heterochiral crystal packing pattern (Fig. 4g). Consequently, (P)-[2b]+ manifested homochiral crystallization, which could inherit (S)-BINAP handedness and higher stability (Fig. S38†). The robust isomerization barriers and well-tailored D–A modules of phospha[5]helicene cations could be realized by introducing strong electron donors at 7,7′/8,8′-positions, and this project was still in progress.
The DFT calculated HOMO of BINAP revealed that electrons were dominated on the P-moieties (Fig. 1a), giving the preferential oxidation of losing one electron in the P-atom and generating the corresponding radical cation 3-[BINAP]+˙ after single electron transfer (SET) oxidation by Mn(III)Cl3. Spin densities analysis demonstrated that P-atoms had the highest spin densities in radical cation 3-[BINAP]+˙ and were preferred sites for the radical attack (Fig. 1a). The calculated Gibbs free energies of the critical transition states also supported the observation that phosphonium attack at the C2′-position was favorable with a low barrier (ΔG‡ = 19.8 kcal mol−1, Fig. 1b), which was reasonable for the observed mild cyclization conditions at near room temperature. With the above results and previous studies on metal-triggered cleavage and annulation of C–P bonds in hand, a plausibly radical phosphonium pathway is depicted in Scheme 3b.16,50–53 The preferential coordination of BINAP with MnCl2 formed an Mn(BINAP)Cl2 intermediate (1a), after which the presence of oxygen caused the oxidation of Mn(II) to Mn(III) complexes (2a). The matched redox potential in the Mn(III) complex led to the formation of radical phosphonium 3-[BINAP]+˙via the SET oxidation between the P-atom and Mn(III)-center. The subsequent attacks with the C2′-position produced the cyclized radical cation 4-[BINAP]+˙, which eliminated the PPh2 residue to form thermodynamically stable [1b]+.54,55
As mentioned above, the installation of the diverse anions at the ionic crystals of phosphahelicens can bring out robust NCI in the packing of those molecules. This might also explain well why our devotion to growing single crystals of other phosphahelicenes (the small halide ion acted as the counterion) ultimately failed, since a more torsional aromatic substituent at the edge-position might significantly result in loose stacking and an unfitting cavity of molecules during the crystallization process. Therefore, the large anion exchange strategy was tried to obtain high-quality single crystals to gain insight into structural information on crystallizing-induced assembly and chiral character. As expected, a series of single crystals were obtained by anion exchange in equivalent halide scavenger (AgBF4 or AgPF6) solutions. [Cl]− anions in enantiomerically enriched [(P)-2b]+[Cl]− and [(M)-2b]+[Cl]− could be exchanged with [BF4]− or [PF6]− anions smoothly, while the crystallographic and stereochemical information was maintained (Fig. S16 and S17†), indicating that the initial chiral bias of cations could be inherited and racemization was hindered during the recrystallization.59,60 In contrast, the prepared [1b]+[BF4]− and [1b]+[PF6]− crystals bearing the PPh2 moieties exhibited expected heterochiral arrangements regardless of stereochemical conformation for precursors (Fig. S14 and S15†). This result demonstrated the racemic nature of the prepared [1b]+[Cl]− under the same conditions, where the isomerization barrier was lower than that of [2b]+[Cl]− by about 5.5 kcal mol−1 (Fig. 4g and S38†). In [(Rac)-3c]+[BF4]−, two edge aromatic rings suffered from folded orientation, huge repulsion, and distortion, thus severely enlarging their dihedral angle for the binaphthyls (θ = 43.6°) and edge phenyls (θ' = 72.9°) as well as the persistence of chiral configuration. In addition, the (M)-isomer or (P)-isomer accumulated into 1-D columnar superhelices along the b-axis with opposite handedness via tight π–π stacking (3.42–3.87 Å) and alternate rotation about 81° and 99° (Fig. 3a and S18g†), where the helical pitch was 23.96 Å. Unexpectedly, homochiral [(M)-7c]+[BF4]− and [(P)-7c]+[BF4]− conglomerates were crystalized in racemic solutions (Fig. 3b), but this chiral self-sorting tendency has not been observed in [(Rac)-3c]+[BF4]−, [(Rac)-1b]+[BF4]−, and [(Rac)-1b]+[PF6]−, manifesting that the chiral self-discrimination was dependent on the subtle variation of cations (Fig. S14, S15, S18, and S19†).60
![]() | ||
| Fig. 3 (a) Crystal structure arrangements of [(Rac)-3c]+[BF4]− and (b) [(M)-7c]+[BF4]−. The hydrogen atoms and anions are omitted for clarity. | ||
Remarkably, all compounds were emissive in chloroform and solid states. [1b]+[Cl]− exhibited structureless green emission at 498 nm with moderate PLQY (Φ = 21%) in chloroform (Fig. 5b). TD-DFT calculations suggested that the fluorescence showed the S1 → S0 transition (Fig. 5b). Moreover, the emission of [1b]+[Cl]− powder displayed a slight bathochromic shift together with a retentive full width at half maximum (FWHM = 76 nm) and an enhanced PLQY (Φ = 34%) owing to the rigidity-enhanced environment in the solid and restriction of intramolecular rotation (RIR) of the tetrahedral C–P scaffolds.68 Furthermore, the fluorescence could be adjusted from deep blue to orange for these serial emitters, including deep blue ([6b]+[Cl]−, 405 nm, Φ = 12%), sky blue ([5b]+[Cl]−, 457 nm, Φ = 77%), yellow ([2c]+[Cl]−, 540 nm, Φ = 39%), orange-yellow ([14b]+[Cl]−, 560 nm, Φ = 35%), and orange ([4c]+[Cl]−, 568 nm in the solid state, Φ = 47%) (Fig. 5c, S41, S43, and Table S12†). Their fluorescence lifetimes (τ) varied from 1.7 to 12.9 ns in chloroform (Fig. 5d). The longer decay of [14b]+[Cl]− and [4c]+[Cl]− than [1b]+[Cl]− could be interpreted as slower relaxation kinetics in the ICT state.67 Notably, all methoxy-substituted compounds at different positions exhibited significant Stokes shifts and higher quantum efficiency, which can be attributed to the enhanced ICT feature of the D–A structures, as evidenced by electronic donation and communication of the bridge methoxy group (Tables S10, S12, and Fig. S32a†).
CPL spectra were also collected for crystalline [(M)/(P)-2b]+[BF4]− powder (Fig. 6 and S44†). Angle-dependent CD/CPL (enantiomerically enriched crystals) and extra CPL (racemic crystals) measurements were performed. CD/CPL signs were unchanged when altering the rotation angles, confirming that the contributions of birefringence and linear dichroism to the CD signals were negligible and the resultant chiral spectra were reliable (Fig. S49†). Enantiomerically enriched [2b]+[BF4]− microcrystals exhibited mirrored CPL emission at 496 nm with moderate luminescence dissymmetry factors (glum) of 3.3 × 10−3/−3.5 × 10−3, which originated from S1 → S0 emission. The experimentally negative CPL signal of [(P)-2b]+[BF4]− was consistent with the TD-DFT simulative results (Fig. S45b, see the ESI† for detailed determination). The |glum(exp)| values and PLQY (41%) for [(P)/(M)-2b]+[BF4]− were comparable to those of reported carbo[n]helicenes and heterohelicenes,69 and the order of magnitude was also equivalent to those of pentavalent phosphahelicene oxides.29,31,66 These results reaffirmed that the synergistic construction of a quaternary C–P bonding environment and helical skeleton was an efficient way to prepare CPL active phosphahelicenes.
Controlling dynamic chirality expression is a formidable challenge and an attractive project for prochiral emitters (their chiral conformation only remained at the condensed phase), for instance, propeller-like TPE derivatives, C3-symmetric dendrimers, axial atropisomers, and macrocycles.70,71 The dynamic chiral conformation in [5]helicenes may be further controlled and induced,38,39 but this conjecture has not been illustrated at present in phosphahelicenes. Herein, an attempt has been made to address this issue through chiral induced LC self-assembly for phospha[5]helicenes. First, we designed and synthesized a pair of axial chiral guests-A ((R)/(S)-BINOL-CN) and guests-B (phospha[5]helicenes) and screened for compatibility and helical twist power (HTP) between the LC host (5CB) and guests (Fig. 7a and b). A ternary cholesteric LC system has been prepared by annealing of the polar 5CB host, racemic phosphorus[5]helicenes (2.0 wt%, [6b]+[Cl]− or [2c]+[Cl]− or [4c]+[Cl]−), and chiral BINOL-CN (0.5–4.0 wt%) mixtures (details are listed in the ESI†). Pure 5CB exhibited a Schlieren texture in the traditional nematic phase at room temperature under a polarized optical microscope (POM), while a uniform fingerprint texture was observed in ternary cholesteric LCs with periodic helicity (Fig. 7c and d). As the concentration of (S)-BINOL-CN increased, the average helical pitch of the fingerprint gradually decreased (Fig. S60 and S61†). Finally, the fingerprint texture transformed into an oily streak texture (4.0 wt%), where the XRD diffraction peak was absent in the small-angle region (Fig. S61e†), confirming that the orientational helical superstructures were induced.72,73
CD spectra of the cholesteric LC film doped with fixed contents of emitters (2.0 wt%) and (S)/(R)-BINOL-CN (4.0 wt%) are shown in Fig. 7e. All prepared (S)-N* and (R)-N* films produced strong negative and positive CD signals at 300–500 nm with large gabs (−0.24 to −0.37, +0.21 to +0.36), respectively. The corresponding CPL spectra of enantiomeric LC films (4c-N*) revealed high glum at 549 nm, up to (+0.51, −0.48), while the enantiomeric 2c-N* and 6b-N* LC films exhibited slightly lower glum at 526 nm (+0.10, −0.14) and 400 nm (+0.22, −0.23), respectively (Fig. 7f). These values were comparable to those of the reported supramolecular helicene polymers and other LC systems.72–75 Furthermore, the CPL emission of phosphahelicenes was blue-shifted about 14–19 nm in LC films compared to their solid fluorescence. The blue-shifted emission was also confirmed in doped PMMA matrices (Fig. S47†), which assuredly revealed a uniform dilution and suppressed π–π stacking of emitters in LCs. Nevertheless, 6b-N* exhibited an integrated deep blue emission due to collective CPL contribution from the host and guests (Fig. S55a and S57a†). Interestingly, the CPL contribution of BINOL-CN has not been observed in 2c-N* and 4c-N* LCs. Although the emission of helicenes can be directly excited at 350 nm, the fluorescence output might be accompanied by partial FRET in this unique system. Both the 5CB host and BINOL-CN donor produced deep blue fluorescence at 382 nm which was predominated by the ICT transition between BINOL and cyanobiphenyls (Fig. S55†), which overlapped with the ICT absorption of the phosphahelicene acceptor (Fig. S56 and S57a†). Hence, we further explored the fluorescence of various molar proportions of [4c]+[Cl]− in host LCs. As molar ratios of helicenes gradually increased, the fluorescence intensity and decay of 5-CB@BINOL-CN appeared to decrease (Fig. S58†). These results provided further evidence for energy transfer from the donor to the acceptor.66,72 In contrast, the guest [6b]+[Cl]− absorption onset peak (370 nm) was out of tune with the host emission region (382 nm), leading to an interrupted energy transfer between the donor and acceptor (Fig. 5a).
To reveal the mechanism of increased CPL emission in N*-LCs, excitation wavelength-dependent CPL spectra of N*-4c were also measured in quartzose cells (Fig. S59†). The CPL spectra of 4c-N* displayed larger glum values in the excitation region of FRET (λem = 350 nm) than direct excitation at 400 nm, demonstrating that chiral FRET could successfully promote the enlargement effect of the induced CPL signal in an intrinsically helical assembly environment.62,76 CPL emissions of emitters were interrelated to chiral BINOL-CN, indicating the formation of chirality transmission and coupling through the π−π, dipole–dipole interactions of the polar aromatic units in the helical host–guest system.76 In addition, no Bragg reflection and structural color were observed in the visible light region due to the sizeable helical pitches, and chirality transmission was interrupted in the phase-separated devices (the N*-LC layer and heliceneslayer were isolated, Fig. S59†), which firmly ruled out the physical reflection of CPL and verified the CPL transmission process.72,77 This suggested that increased CPL of phosphahelicenes benefited from a spiral arrangement and chiral energy transmission.
Footnote |
| † Electronic supplementary information (ESI) available. CCDC 2130476, 2130479, 2130454, 2130459, 2225461, 2225466–2225474, 2225477–2225478, and 2225482. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3sc03201a |
| This journal is © The Royal Society of Chemistry 2023 |