Open Access Article
Jieli
Lin
a,
Shihua
Liu
a,
Jie
Zhang
a,
Hansjörg
Grützmacher
ab,
Cheng-Yong
Su
a and
Zhongshu
Li
*a
aLIFM, IGCME, School of Chemistry, Sun Yat-Sen University, Guangzhou 510006, China. E-mail: lizhsh6@mail.sysu.edu.cn
bDepartment of Chemistry and Applied Biosciences, ETH Zürich, Vladimir-Prelog-Weg 1, Zürich 8093, Switzerland
First published on 11th September 2023
E,Z-isomers display distinct physical properties and chemical reactivities. However, investigations on heavy main group elements remain limited. In this work, we present the isolation and X-ray crystallographic characterization of N-heterocyclic vinyl (NHV) substituted diphosphenes as both E- and Z-isomers (L
CH–P
P–CH
L, E,Z-2b; L = N-heterocyclic carbene). E-2b is thermodynamically more stable and undergoes reversible photo-stimulated isomerization to Z-2b. The less stable Z-isomer Z-2b can be thermally reverted to E-2b. Theoretical studies support the view that this E ↔ Z isomerization proceeds via P
P bond rotation, reminiscent of the isomerization observed in alkenes. Furthermore, both E,Z-2b coordinate to an AuCl fragment affording the complex [AuCl(η2-Z-2b)] with the diphosphene ligand in Z-conformation, exclusively. In contrast, E,Z-2b undergo [2 + 4] and [2 + 1] cycloadditions with dienes or diazo compounds, respectively, yielding identical cycloaddition products in which the phosphorus bound NHV groups are in trans-position to each other. DFT calculations provide insight into the E/Z-isomerisation and stereoselective formation of Au(I) complexes and cycloaddition products.
SiMes2 (ref. 1), and a diphosphene, Mes*P
PMes*2 (A, Fig. 1a), in 1981.3 The species containing heavy main group elements often exhibit unique structures and reactivity patterns compared to their lighter congeners. Some are characterized by “trans-bent” E
E double bonds (that is, the molecule R2E
ER2 is not planar but shows ER2 interplanar angles >0°) and show unique reactivity.3a–e,g–i The kinetic barriers for oligo/polymerization are much lower than for the lighter congeners with E = element from the second period, and the successful isolation of compounds with E from the higher periods requires the utilization of sterically encumbering substituents to introduce kinetic protection. Consequently, cis-folded isomers of R2E
ER2 or Z-isomers of RE
ER with heavy main group elements are rare due to increased steric repulsion caused by these substituents.3a,b,d,4
![]() | ||
| Fig. 1 Schematic representation of (a) irradiation of isolable E-diphosphene; (b) decomposition of isolable Z-diphosphenes; and (c) reversible E ↔ Z isomerization of isolable diphosphenes. | ||
Diphosphenes are particularly intriguing; due to the diagonal relationship in the periodic table between carbon and phosphorus they behave to a certain extent like olefins.3b,d,5 Numerous E-diphosphenes were fully characterized, but the synthesis and isolation of Z-diphosphenes remains challenging.4e,6 The first spectroscopic detection of transient Z-diphosphene, Z-A, was found by laser irradiation of E-A at −80 °C. Z-A would revert back to the thermodynamically more stable E-A upon warming the solution to room temperature (RT) (Fig. 1a).6b,c Remarkably, Niecke et al. have isolated and crystallographically characterized a series of Z-diphosphenes (Z-RP
PMes*, R = tBu(H)N, adamantyl(H)N, Et3C(H)N, 2,4,6-iPrC6H2(H)N, (Me3Si)2N(SiMe3)N, B, Fig. 1b).7 Among these, one diphosphene RP
PMes* with a sterically very demanding amido substituent [R = (Me3Si)2N(SiMe3)N] could be isolated as both the E- and Z-isomers. These compounds displayed E,Z-isomerization in solution at RT.7b To the best of our knowledge, this is the only example where both the E- and Z-isomers of identical molecular composition could be isolated, although these amido-substituted diphosphenes undergo slow metatheses reactions at RT yielding symmetrical diphosphenes Mes*P
PMes* and RP
PR, likely via cyclotetraphosphanes formed in head-to-head dimerizations as intermediates (Fig. 1b). More recently, both diphosphasilene6a and diphosphene8 substituted Z-diphosphene, [Si]
P–P
P–P
[Si] ([Si] = [PhC(NtBu)2SiN(SiMe3)2]) and L
P–P
P–P
L (L = N-heterocyclic carbene (NHC)), were reported but likewise show limited stability.
Nevertheless, the presence of the RR′N–P
P or R
P–P
P–P
R fragments in Z-diphosphenes6a,7 indicates some structural and electronic features for constructing stable Z-diphosphenes: (i) the sterically protecting group points away from the central P
P group to impose sufficient kinetic stability while at the same time the steric repulsion between the two substituents in Z-conformation is minimized. (ii) Electron delocalization as provided by co-planar moderately π-electron donating groups across the P
P bond should increase the thermodynamic stability of both the E- and Z-isomers. We therefore reasoned that N-heterocyclic vinyl (NHV = L
CR) substituted diphosphenes of the type L
CR–P
P–CR
L could likewise be promising candidates for the observation and eventual isolation of persistent Z-diphosphenes. Both the groups of Stephan and Ghadwal independently reported NHV-substituted E-diphosphenes L
C(R)–P
P–C(R)
L, E-C (L = cyclic alkyl amino carbene (CAAC), R = tBu, adamantyl)9 and E-D (L = NHC, R = phenyl)10 (Fig. 1c). In these compounds, the presence of bulky R substituents on the NHV groups prevent the observation of the corresponding Z-isomers. In this study, we have successfully isolated and characterized diphosphene L
C(R)–P
P–C(R)
L 2b in both E- and Z-configurations using the smallest possible substituent R = H. Specifically, E-2b can be converted reversibly to Z-2b by a photo-stimulated process while Z-2b is reverted to E-2b by a thermal process. Notably, only a slight increase in bulkiness in passing from R = H to Me allows only the E-isomer E-2a to be detected (Fig. 1c). The thermodynamic and kinetic parameters for E/Z-isomerisation as well as differences in coordination and cyclo-addition reactions between E- and Z-2b could be studied.
C(R)–PCl2, L = SIPr = 1,3-bis-(2,6-diisopropylphenyl)imidazoline-2-ylidine) were prepared as white powders [1a: δ(31P) = 191.9 ppm; 1b: δ(31P) = 183.7 ppm; see the ESI† for further details].10,11 Treatment of 1a with excess Mg powder in tetrahydrofuran (THF) at RT under vigorous stirring overnight in the dark afforded exclusively E-2a (L
C(Me)–P
P–C(Me)
L), which is photolytically and thermally stable. After work-up, E-2a was obtained in pure form as a red powder in 78% yield. Reaction of 1b with Mg powder under the same conditions afforded a mixture of E-2b and Z-2b (L
C(H)–P
P–C(H)
L) in a ratio of 1.0
:
0.3 (Fig. 2a). After evaporation of THF, the solid residue was treated with toluene in order to extract the mixture of E/Z-2b from MgCl2. Heating this solution of E/Z-2b at 110 °C for 20 minutes in the dark led to an enrichment of E-2b (E-2b
:
Z-2b = 1.0
:
0.1). In contrast, irradiation of the toluene solution of the mixture of E/Z-2b with UV light (Hg lamp) or LED light (520 nm) at 0 °C for 30 minutes led to the predominant formation of Z-2b (E-2b
:
Z-2b = 1.0
:
21.6). Evaporation of solvent toluene from these reaction mixtures after irradiation or heating, and recrystallization of the solid residues from cold pentane at −30 °C yielded a highly air-sensitive red crystalline powder of E-2b or a crystalline orange powder of Z-2b in 62% and 67% yields, respectively.
![]() | ||
| Fig. 2 (a) Synthesis of compounds 2a and 2b; (b) the E → Z isomerization of E-2b in the solid state as a thin layer in between two glass plates. | ||
Multi-nuclear NMR spectra, single crystal X-ray diffraction (XRD), and high-resolution mass spectrometry were applied to characterize and confirm the molecular structures of E-2a, E-2b, and Z-2b. Characteristic singlet resonances at δ(31P) = 382.6, 379.6, or 259.5 ppm were observed in the 31P NMR spectra of E-2a, E-2b, and Z-2b, respectively. These 31P NMR shifts at high frequencies are within the range observed for other diphosphenes as well. In agreement with literature reported data, E-configured diphosphenes show resonances by about 100 ppm at higher frequencies compared to those of the Z-isomers.6g,12 In the 1H NMR spectra, singlet resonances at δ(1H) = 1.85 (E-2a PCCH3), δ(1H) = 4.98 (E-2b PCH), and δ(1H) = 4.61 ppm (Z-2b PCH) are attributed to the protons of the R substituents at the NHV moieties, respectively.
The stereo-configurations of 2a and 2b are confirmed unambiguously by XRD methods (Fig. 3). The P
P bond lengths in the E-configured isomers E-2a [2.052(7) Å] and E-2b [2.0685(12) Å], and the one in the Z-isomer Z-2b [2.058(6) Å] are remarkably similar and typical of P
P double bonds [Σrcov(P–P) = 2.22 Å, Σrcov(P
P) = 2.04 Å].13 The P–C and C
C bonds [1.765(3)–1.7863(13) Å and 1.362(4)–1.375(2) Å] vary only slightly in 2a and 2b and are typical of π-conjugated delocalized P
C or C
C bonds [Σrcov(P–C) = 1.86 Å, Σrcov(P
C) = 1.69 Å, Σrcov(C–C) = 1.50 Å, Σrcov(C
C) = 1.34 Å].13a,b As expected, the C–P
P–C fragments are nearly planar in E-2a (∠CPPC = 0.0°) and E-2b (∠CPPC = 0.2°), but show a small deviation from planarity in Z-2b (∠CPPC = 13.3°) indicating a higher steric congestion in the Z-isomer.
Toluene solutions of the E-configured isomers E-2a and E-2b are deep red and show strong absorptions at λmax = 518 (1.8 × 104 M−1 cm−1) and 534 nm (1.8 × 104 M−1 cm−1). The Z-isomer Z-2b exhibits a strong absorption at lower wavenumbers with λmax = 482 nm (2.3 × 104 M−1 cm−1). In combination with TD-DFT calculations (Fig. S73–S78†), these absorptions are mainly attributed to the allowed π–π* transitions (HOMO → LUMO). The larger HOMO–LUMO gap of the Z-isomer is due to a lower HOMO energy (Z-2b: −5.27 eV vs. E-2b: −5.06 eV) and a higher LUMO energy (Z-2b: −0.44 eV vs. E-2b: −0.51 eV). In addition, all compounds show a very weak absorption at shorter wavelengths, namely 416 (2.5 × 103 M−1 cm−1), 401 (3.0 × 103 M−1 cm−1), and 359 (2.3 × 103 M−1 cm−1) nm for E-2a, E-2b, and Z-2b, respectively (Fig. 4a). These are assigned to the forbidden n–π* transitions (HOMO−1 → LUMO). Comparable results have been reported for the structurally similar D (Fig. 1d).10 Note, however, that in aryl substituted diphosphenes such as Mes*P
PMes* the ordering of these transitions is inverted and the n–π* is at higher wavenumbers while the π–π* occurs at shorter wavenumbers.14
![]() | ||
| Fig. 4 UV-vis absorption spectra for toluene solution of E-2a (dashed red), E-2b (red) and Z-2b (black). | ||
The kinetics for the E ↔ Z isomerization of 2b was probed by 1H NMR spectroscopy using hexamethylenetetramine as an internal standard. The pseudo-first-order rate constants (k) were calculated based on the consumption of the E-2b or Z-2b for photo- or thermally stimulated E,Z-isomerization. The plots of ln[E-2b] vs. t (Fig. 5a) and ln[Z-2b] vs. t (Fig. S59† and 5b) indicate first order kinetics with k520nm = 4.4 × 10−3 s−1, k455nm = 1.0 × 10−2 s−1, and k348.15K = 5.6 × 10−4 s−1, respectively. The photostationary (PSS) equilibria were established within 300 s or 750 s for the E-2b → Z-2b or E-2b ← Z-2b isomerization process yielding a mixture of E-2b
:
Z-2b = 0.19 or 4.40 under irradiation at 520 or 455 nm, respectively. The relative molar extinction coefficient of E,Z-2b at 520 or 455 nm was determined to be Z-2b
:
E-2b = 0.18 or 4.76 (see Fig. S50 and S51† for details). These results fit well to eqn (1) for the PSS ensuring equal quantum yields (Φ) for the photoisomerization reactions in both directions.19
![]() | (1) |
Furthermore, the temperature dependence of the isomerization rate constructed from the kinetic data measured at 5 K intervals from 348.15 K to 368.15 K gives the experimental activation energy for the thermal Z-2b → E-2b conversion as Ea = 23.4 ± 0.9 kcal mol−1. The pre-exponential factor can be expressed at these temperatures by the Arrhenius equation k = 2.72 × 1011 × e−23.4/RT (Fig. 5c). This high pre-exponential factor precludes the population of excited triplet states in the thermal Z-2b → E-2b isomerization process.20 The Eyring plot constructed from these data affords the experimental activation parameters as ΔH‡ = 22.7 ± 0.9 kcal mol−1, ΔS‡ = −8.6 ± 2.4 e.u., and ΔG‡(298.15K) = 25.3 ± 1.6 kcal mol−1 (Fig. S64†). The activation barrier of the thermal isomerization is comparable to those of the previously reported diphosphenes Mes*P
PMes* [ΔG‡(273K) = 20.3 kcal mol−1]6b and RP
PMes* (ΔG‡(293K) = 25.5 kcal mol−1, R = (Me3Si)2N(Me3Si)N, Fig. 1b).7b For comparison, the activation barriers for stilbene21 or azobenzene22 are about 40 or 23 kcal mol−1, respectively. Moreover, a van't Hoff analysis (ΔH = 1.6 ± 0.1 kcal mol−1, ΔS = 7.6 ± 0.2 e.u.) revealed that the free energy of Z-2b is slightly lower, ΔG(298.15K) = −0.6 ± 0.1 kcal mol−1, than that of E-2b (Fig. 5d). In conclusion, these special factors (small energy difference between E- and Z-2b and relatively high activation barrier for the thermal Z-2b → E-2b isomerization process) allow the isolation of diphosphene 2b in both E- and Z-configurations.
To provide a deeper understanding of the E,Z-isomerization mechanism, density functional theory (DFT) calculations were carried out at the M062X-D3/Def2TZVP-SMD(toluene)//M062X-D3/Def2SVP level of theory.23 Apart from E,Z-2b we included also the methyl-substituted derivates E,Z-2a in this study although an E/Z-isomerism could not be observed experimentally. Possible minimum energy reaction pathways (MERPs) for both the photolytic E-X ↔ Z-X and thermal Z-X → E-X (X = 2a, 2b) are shown in Fig. 6. Upon irradiation, E-2a and E-2b are excited to the triplet states 3T-2a (26.7 kcal mol−1) and 3T-2b (23.6 kcal mol−1).24 The structures of the triplet states 3T-2a and 3T-2b are very similar and notably contain a significantly elongated P–P bond (2.21 Å), which is in the range of a single bond, and C–P–P–C dihedral angles of 106.0°. These structures are very different compared to the singlet ground state structures of E-2a and E-2b (P
P: 2.05 Å; C–P–P–C = 180.0°). Subsequently, these excited triplet states relax via a spin flipping process to the E- or Z-isomers in their singlet ground states in exergonic reactions. In case of 2a, the resulting Z-2a will quickly isomerize to the thermodynamically more stable E-2a (−11.2 kcal mol−1) via a relatively small activation barrier TSR-2a (13.9 kcal mol−1). In case of 2b, the E/Z ratio of the resulting mixture depends on the relative molar extinction coefficient of E,Z-2b. At elevated temperatures, Z-2b isomerizes via the activated complexes TSR-2b (22.7 kcal mol−1), which has a singlet electronic configuration, affording the thermodynamically more stable E-isomers E-b (−3.3 kcal mol−1). The calculations are in good agreement with the experimental observations: Z-2a cannot be detected because of its thermodynamic instability versus E-2a (+11.2 kcal mol−1) and low activation barrier for the thermal isomerization (Ea = 13.9 kcal mol−1). On the other hand, Z-2b can be isolated because it is almost isoenergetic with E-2b and Ea for the thermal isomerization is much higher (25.3 ± 1.6 kcal mol−1). Note that the calculated activation barriers predicted at the same level of theory for the Z → E isomerization process of the olefin stilbene [TSR = 43.6 kcal mol−1vs. 40 kcal mol−1 (exp.)]21 is much higher while the one of azobenzene [TSI = 25.5 kcal mol−1vs. 23 kcal mol−1 (exp.)]22 is in the same range (Fig. S79†). For comparison we also investigated possible E/Z-isomerization pathways for the simple divinyl substituted compound CH2
CH–P
P–CH
CH2 and found that the activation barrier EaR = 24.3 kcal mol−1 for a rotation around the P,P-bond is comparable to the one in E-2b (the activation barrier for inversion at one phosphorus center – an alternative process for E/Z-isomerization – is much higher in energy at EaI = 50.1 kcal mol−1) (Fig. S80†). Similar data have been theoretically predicted for the Z → E isomerization of HP
PH (Fig. S81†).25
The molecular structure of [AuCl(2b)] was determined by XRD methods (Fig. 7b) and shows that 2b is (i) η2-bound to the AuCl moiety and (ii) the substituents adopt a cis-configuration with respect to the P–P vector indicating that, remarkably and unexpectedly, the reaction of E-2b with [AuCl(SMe2)] induces E-2b → Z-2b isomerization. In the complex [AuCl(η2-Z-2b)], the P1–P2 bond length is elongated to 2.1254(12) Å with respect to uncoordinated Z-2b [2.0580(6) Å]. This observation indicates significant electron donation from the d10-valence electron configured Au(I) ion into the π*(P
P) anti-bonding orbital. The Au–P1 and Au–P2 bonds are almost identical in length [P1–Au1: 2.4580(11) Å; P2–Au1: 2.4280(11) Å]. The metal center resides in a trigonal planar geometry [sum of bond angles Σ°(Au) = 359.8°]. The dihedral angle between the Au–P1–P2 and C1–P1–P2–C3 planes is 73.9°. Hand in hand with the elongation of the coordinated P
P bond goes a reduction of the ∠CPPC torsion angle from 13.3° (Z-2b) to 5.7° in the complex [AuCl(η2-Z-2b)] indicating reduced steric repulsion between two NHV groups. To the best of our knowledge, [AuCl(η2-Z-2b)] represents the first crystallographically characterized mononuclear η2-Z-diphosphene complex and the only η2-bound diphosphene Au(I) complex isolated so far (vide infra).7,18,26
DFT calculations were carried out in order to gain deeper insight into the selective formation of [AuCl(η2-Z-2b)]. Both the full model and a simplified model complex [AuCl(η2-Z-2bH)] where the 2,6-diisopropylphenyl (Dipp) groups were replaced with H atoms were investigated at the BP86-D3BJ/Def2TZVP-SMD(THF)//BP86-D3BJ/Def2SVP level of theory (Fig. 8).23 For the full model, the stability of four different conformers follows the order [AuCl(η2-Z-2b)] (−1.8 kcal mol−1) > [AuCl(η1-E-2b)] (0 kcal mol−1) > [AuCl(η2-E-2b)] (0.1 kcal mol−1) > [AuCl(η1-Z-2b)] (4.4 kcal mol−1). The highest activation barrier of 22.3 kcal mol−1 (TS2-2b) is attributed to the rotation of the P
P bond during the [AuCl(η1-E-2b)] → [AuCl(η1-Z-2b)] isomerization process, which is only slightly smaller than the calculated barrier for the E/Z-isomerization of the uncoordinated diphosphene. On the other hand, the conversion from η1- to η2-gold complexes for both E- and Z-isomers proceeds smoothly viaTS1-2b (7.4 kcal mol−1) and TS3-2b (0.6 kcal mol−1). For the simplified model, similar energy barriers are obtained for the intramolecular conversion among these four conformers. However, the stability order differs and follows the order [AuCl(η1-E-2bH)] (0 kcal mol−1) > [AuCl(η1-Z-2bH)] (0.3 kcal mol−1) > [AuCl(η2-Z-2bH)] (0.6 kcal mol−1) > [AuCl(η2-E-2bH)] (2.0 kcal mol−1). The higher stability (although marginal) of the η1-bound complexes agrees with the one previously reported for three Au(I) bisarene-substituted E-diphosphene complexes, in which the P
P bond is shortened with respect to the uncoordinated diphosphene.26 Note that diphosphenes in E-configuration are frequently bound in η2-fashion to other transition metal centers.27
![]() | ||
| Fig. 8 MERP for the intramolecular conversion between η1- and η2-gold complexes with both models 2b and 2bH. | ||
Again, DFT calculations were performed to provide a better understanding of the reactivity difference in between E-2b and Z-2b. The reaction mechanism for the [2 + 4] cycloaddition reaction between DMBD and model 2b or 2bH was studied at the M062X-D3/Def2TZVP-SMD(toluene)//M062X-D3/Def2SVP level of theory (Fig. 11).23 In both cases, the [2 + 4] cycloaddition reactions proceed in a concerted fashion via one activated complex at the corresponding transition states. In the reaction between DMBD and the simplified E-configured model E-2bH the transition state TS6H (15.2 kcal mol−1) is about 5 kcal mol−1 lower than the one for the reaction with Z-2bH (TS4H = 20.3 kcal mol−1 with respect to Z-2bH). Both products E-3H (−17.0 kcal mol−1) and Z-5H (−12.2 kcal mol−1) are formed in exergonic reactions, which differ by 4.8 kcal mol−1 in favour of the E-isomer. The kinetic and thermodynamic formation of the E-configured isomer becomes even more favourable in the reactions with the sterically more hindered E/Z-2b as diene acceptors. Here TS6 on the MERP in the reaction with E-2b is lower by 9.0 kcal mol−1 than TS4 (on the MERP with Z-2b) and the E-configured product E-3 is more stable by 14.6 kcal mol−1 when compared to Z-3. The latter is formed in an almost thermoneutral reaction (ΔGr = −3.6 kcal mol−1) between Z-2b and DMBD while the formation of the E-isomer E-3 is exergonic by −14.9 kcal mol−1. Not unexpected, with increasing steric encumbrance the transition state energies increase while the product stabilities decrease. But the calculated barriers for the [2 + 4] cycloaddition between E/Z-2b and DMBD are still relatively low and comparable to the barrier for E → Z isomerization to allow a reaction at room temperature, which ultimately leads selectively to the thermodynamically more stable product E-3. Thus, the calculations agree well with the experimental observations.
P double bond of about the same length and confirm their stereo configuration. Kinetic studies reveal the first-order process for both photo- and thermally stimulated E ↔ Z isomerization. In combination, the experimental and theoretical data show that the E ↔ Z isomerization of diphosphene proceeds via rotation around the P
P double bond. However, a similar E ↔ Z isomerization has not been reported for the nitrogen analogue of E-2b.30 Both E,Z-2b react with the AuCl fragment to afford exclusively the thermodynamically slightly more favoured complex [AuCl(η2-Z-2b)], which is the first mononuclear metal complex of a diphosphene in Z-conformation. The activation barrier for E → Z-isomerisation of the η1-bound AuCl complexes, which are intermediates on the MERP that give [AuCl(η2-Z-2b)], is slightly smaller (≈4 kcal mol−1) than in the uncoordinated diphosphene. It is presently not clear why in this particular case the Au(I) complex with Z-2b is more stable. The opposite is observed when a mixture of E/Z-2b is employed in a cycloaddition with a diene or diazo compound. Here, the Z-isomer is unreactive and must isomerize slowly to E-2b in order to allow a reaction. Compound E-2b readily engages in a [2 + 4] or [2 + 1] cycloaddition reaction and as a consequence, exclusively the cycloaddition product with an E-arrangement of the substituents with respect to the P–P vector is obtained.
Footnote |
| † Electronic supplementary information (ESI) available: Synthesis and characterization of compounds, NMR spectra, and crystallographic and computational details. CCDC 2245822–2245824 and 2270969–2270971. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3sc04506d |
| This journal is © The Royal Society of Chemistry 2023 |