首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Efficient photoswitching in the solid-state remains rare, yet is highly desirable for the design of functional solid materials. In particular, for molecular solar thermal energy storage materials high conversion to the metastable isomer is crucial to achieve high energy density. Herein, we report that 4-methoxyazobenzene (MOAB) can be occluded into the pores of a metal–organic framework Zn2(BDC)2(DABCO), where BDC = 1,4-benzenedicarboxylate and DABCO = 1,4-diazabicyclo[2.2.2]octane. The occluded MOAB guest molecules show near-quantitative EZ photoisomerization under irradiation with 365 nm light. The energy stored within the metastable Z-MOAB molecules can be retrieved as heat during thermally-driven relaxation to the ground-state E-isomer. The energy density of the composite is 101 J g−1 and the half-life of the Z-isomer is 6 days when stored in the dark at ambient temperature.

4-Methoxyazobenzene can be occluded into the pores of a MOF and show near-quantitative EZ photoisomerization under irradiation with 365 nm light. The energy density of the composite is 101 J g−1 and the half-life of the Z-isomer is 6 days.  相似文献   

2.
The chromophore of the green fluorescent protein (GFP) is critical for probing environmental influences on fluorescent protein behavior. Using the aqueous system as a bridge between the unconfined vacuum system and a constricting protein scaffold, we investigate the steric and electronic effects of the environment on the photodynamical behavior of the chromophore. Specifically, we apply ab initio multiple spawning to simulate five picoseconds of nonadiabatic dynamics after photoexcitation, resolving the excited-state pathways responsible for internal conversion in the aqueous chromophore. We identify an ultrafast pathway that proceeds through a short-lived (sub-picosecond) imidazolinone-twisted (I-twisted) species and a slower (several picoseconds) channel that proceeds through a long-lived phenolate-twisted (P-twisted) intermediate. The molecule navigates the non-equilibrium energy landscape via an aborted hula-twist-like motion toward the one-bond-flip dominated conical intersection seams, as opposed to following the pure one-bond-flip paths proposed by the excited-state equilibrium picture. We interpret our simulations in the context of time-resolved fluorescence experiments, which use short- and long-time components to describe the fluorescence decay of the aqueous GFP chromophore. Our results suggest that the longer time component is caused by an energetically uphill approach to the P-twisted intersection seam rather than an excited-state barrier to reach the twisted intramolecular charge-transfer species. Irrespective of the location of the nonadiabatic population events, the twisted intersection seams are inefficient at facilitating isomerization in aqueous solution. The disordered and homogeneous nature of the aqueous solvent environment facilitates non-selective stabilization with respect to I- and P-twisted species, offering an important foundation for understanding the consequences of selective stabilization in heterogeneous and rigid protein environments.

Simulations on the aqueous green fluorescent protein (GFP) chromophore (in the equilibrium and non-equilibrium regimes) reveal that observed biexponential fluorescence originates from two competing torsional deactivation pathways.  相似文献   

3.
Natural and laboratory-guided evolution has created a rich diversity of fluorescent protein (FP)-based sensors for chloride (Cl). To date, such sensors have been limited to the Aequorea victoria green fluorescent protein (avGFP) family, and fusions with other FPs have unlocked ratiometric imaging applications. Recently, we identified the yellow fluorescent protein from jellyfish Phialidium sp. (phiYFP) as a fluorescent turn-on, self-ratiometric Cl sensor. To elucidate its working mechanism as a rare example of a single FP with this capability, we tracked the excited-state dynamics of phiYFP using femtosecond transient absorption (fs-TA) spectroscopy and target analysis. The photoexcited neutral chromophore undergoes bifurcated pathways with the twisting-motion-induced nonradiative decay and barrierless excited-state proton transfer. The latter pathway yields a weakly fluorescent anionic intermediate , followed by the formation of a red-shifted fluorescent state that enables the ratiometric response on the tens of picoseconds timescale. The redshift results from the optimized π–π stacking between chromophore Y66 and nearby Y203, an ultrafast molecular event. The anion binding leads to an increase of the chromophore pKa and ESPT population, and the hindrance of conversion. The interplay between these two effects determines the turn-on fluorescence response to halides such as Cl but turn-off response to other anions such as nitrate as governed by different binding affinities. These deep mechanistic insights lay the foundation for guiding the targeted engineering of phiYFP and its derivatives for ratiometric imaging of cellular chloride with high selectivity.

We discovered an interplay between proton transfer and conformational reorganization that powers a standalone fluorescent-protein-based excitation-ratiometric biosensor for chloride imaging.  相似文献   

4.
Multiply-bonded main group metal compounds are of interest as a new class of reactive species able to activate and functionalize a wide range of substrates. The aluminium sulfido compound K[Al(NONDipp)(S)] (NONDipp = [O(SiMe2NDipp)2]2−, Dipp = 2,6-iPr2C6H3), completing the series of [Al(NONDipp)(E)] anions containing Al–E{16} multiple bonds (E{16} = O, S, Se, Te), was accessed via desulfurisation of K[Al(NONDipp)(S4)] using triphenylphosphane. The crystal structure showed a tetrameric aggregate joined by multiple K⋯S and K⋯π(arene) interactions that were disrupted by the addition of 2.2.2-cryptand to form the separated ion pair, [K(2.2.2-crypt)][Al(NONDipp)(S)]. Analysis of the anion using density functional theory (DFT) confirmed multiple-bond character in the Al–S group. The reaction of the sulfido and selenido anions K[Al(NONDipp)(E)] (E = S, Se) with CO2 afforded K[Al(NONDipp)(κ2E,O-EC{O}O)] containing the thio- and seleno-carbonate groups respectively, consistent with a [2 + 2]-cycloaddition reaction and C–E bond formation. An analogous cycloaddition reaction took place with benzophenone affording compounds containing the diphenylsulfido- and diphenylselenido-methanolate ligands, [κ2E,O-EC{O}Ph2]2−. In contrast, when K[Al(NONDipp)(E)] (E = S, Se) was reacted with benzaldehyde, two equivalents of substrate were incorporated into the product accompanied by formation of a second C–E bond and complete cleavage of the Al–E{16} bonds. The products contained the hitherto unknown κ2O,O-thio- and κ2O,O-seleno-bis(phenylmethanolate) ligands, which were exclusively isolated as the cis-stereoisomers. The mechanisms of these cycloaddition reactions were investigated using DFT methods.

Reaction of Al–E (E = S, Se) multiple bonds with C Created by potrace 1.16, written by Peter Selinger 2001-2019 O functionalities generates new C–E bonds.  相似文献   

5.
Molecular antiferroelectrics (AFEs) have taken a booming position in the miniaturization of energy storage devices due to their low critical electric fields. However, regarding intrinsic competitions between dipolar interaction and steric hindrance, it is a challenge to exploit room-temperature molecular AFEs with high energy storage efficiency. Here, we present a new 2D hybrid perovskite-type AFE, (i-BA)2(FA)Pb2Br7 (1), which shows ultrahigh energy storage efficiencies at room temperature. Most strikingly, the typical double PE hysteresis loops afford an ultrahigh storage efficiency up to ∼91% at low critical electric fields (Ecr = 41 kV cm−1); this Ecr value is much lower than those of state-of-the-art AFE oxides, revealing the potential of 1 for miniaturized energy-storage devices. In terms of the energy storage mechanism, the dynamic ordering and antiparallel reorientation of organic cations trigger its AFE-type phase transition at 303 K, thus giving a large spontaneous electric polarization of ∼3.7 μC cm−2, while the increasement of steric hindrance of the organic branched-chain i-BA+ spacer cations stabilizes its antipolar sublattices. To the best of our knowledge, this exploration of achieving ultrahigh energy storage efficiency at such a low critical electric field is unprecedented in the AFE family, which paves a pathway for miniaturized energy storage applications.

We report a room-temperature molecular AFE, (i-BA)2(FA)Pb2Br7, which shows ultrahigh η of ∼91% at quite low electric fields. This finding will be advantageous to the progress of high energy-storage applications.  相似文献   

6.
The occurrence of planar hexacoordination is very rare in main group elements. We report here a class of clusters containing a planar hexacoordinate silicon (phSi) atom with the formula SiSb3M3+ (M = Ca, Sr, Ba), which have D3h (1A1′) symmetry in their global minimum structure. The unique ability of heavier alkaline-earth atoms to use their vacant d atomic orbitals in bonding effectively stabilizes the peripheral ring and is responsible for covalent interaction with the Si center. Although the interaction between Si and Sb is significantly stronger than the Si–M one, sizable stabilization energies (−27.4 to −35.4 kcal mol−1) also originated from the combined electrostatic and covalent attraction between Si and M centers. The lighter homologues, SiE3M3+ (E = N, P, As; M = Ca, Sr, Ba) clusters, also possess similar D3h symmetric structures as the global minima. However, the repulsive electrostatic interaction between Si and M dominates over covalent attraction making the Si–M contacts repulsive in nature. Most interestingly, the planarity of the phSi core and the attractive nature of all the six contacts of phSi are maintained in N-heterocyclic carbene (NHC) and benzene (Bz) bound SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes. Therefore, bare and ligand-protected SiSb3M3+ clusters are suitable candidates for gas-phase detection and large-scale synthesis, respectively.

The global minimum of SiSb3M3+ (M = Ca, Sr, Ba) is a D3h symmetric structure containing an elusive planar hexacoordinate silicon (phSi) atom. Most importantly, the phSi core remains intact in ligand protected environment as well.

Exploring the bonding capacity of main-group elements (such as carbon or silicon) beyond the traditional tetrahedral concept has been a fascinating subject in chemistry for five decades. The 1970 pioneering work of Hoffmann and coworkers1 initiated the field of planar tetracoordinate carbons (ptCs), or more generally, planar hypercoordinate carbons. The past 50 years have witnessed the design and characterization of an array of ptC and planar pentacoordinate carbon (ppC) species.2–14 However, it turned out to be rather challenging to go beyond ptC and ppC systems. The celebrated CB62− cluster and relevant species15,16 were merely model systems because C avoids planar hypercoordination in such systems.17,18 In 2012, the first genuine global minimum D3h CO3Li3+ cluster was reported to have six interactions with carbon in planar form, although electrostatic repulsion between positively charged phC and Li centers and the absence of any significant orbital interaction between them make this hexacoordinate assignment questionable.19 It was only very recently that a series of planar hexacoordinate carbon (phC) species, CE3M3+ (E = S–Te; M = Li–Cs), were designed computationally by the groups of Tiznado and Merino (Fig. 1; left panel),20 in which there exist pure electrostatic interactions between the negative Cδ− center and positive Mδ+ ligands. These phC clusters were achieved following the so-called “proper polarization of ligand” strategy.Open in a separate windowFig. 1The pictorial depiction of previously reported phC CE3M3+ (E = S–Te; M = Li–Cs) clusters and the present SiE3M3+ (E = S–Te and N–Sb; M = Li–Cs and Ca–Ba) clusters. Herein the solid and dashed lines represent covalent and ionic bonding, respectively. The opposite double arrows illustrate electrostatic repulsion.The concept of planar hypercoordinate carbons has been naturally extended to their next heavier congener, silicon-based systems. Although the steric repulsion between ligands decreases due to the larger size, the strength of π- and σ-bonding between the central atom and peripheral ligands dramatically decreases, which is crucial for stability. Planar tetracoordinate silicon (ptSi) was first experimentally observed in a pentaatomic C2v SiAl4 cluster by Wang and coworkers in 2000.21 Very recently, this topic got a huge boost by the room-temperature, large-scale syntheses of complexes containing a ptSi unit.22 A recent computational study also predicted the global minimum of SiMg4Y (Y = In, Tl) and SiMg3In2 to have unprecendented planar pentacoordinate Si (ppSi) units.23 Planar hexacoordinate Si (phSi) systems seem to be even more difficult to stabilize. Previously, a C2v symmetric Cu6H6Si cluster was predicted as the true minimum,24 albeit its potential energy surface was not fully explored. A kinetically viable phSi SiAl3Mg3H2+ cluster cation was also predicted.25 However, these phSi systems24,25 are only local minima and not likely to be observed experimentally. In 2018, the group of Chen identified the Ca4Si22− building block containing a ppSi center and constructed an infinite CaSi monolayer, which is essentially a two-dimensional lattice of the Ca4Si2 motif.26 Thus, it is still an open question to achieve a phSi atom to date.Herein we have tried to find the correct combination towards a phSi system as the most stable isomer. Gratifyingly, we found a series of clusters, SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba), having planar D3h symmetry with Si at the center of the six membered ring, as true global minimum forms. Si–E bonds are very strong in all the clusters, and alkaline-earth metals interact with the Si center by employing their d orbitals. However, electrostatic repulsion originated from the positively charged Si and M centers for E = N, P, and As dominates over attractive covalent interaction, making individual Si–M contacts repulsive in nature. This makes the assignment of SiE3M3+ (E = N, P, As; M = Ca, Sr, Ba) as genuine phSi somewhat skeptical. SiSb3M3+ (M = Ca, Sr, Ba) clusters are the sole candidates which possess genuine phSi centers as both electrostatic and covalent interactions in Si–M bonds are attractive. The d orbitals of M ligands play a crucial role in stabilizing the ligand framework and forming covalent bonds with phSi. Such planar hypercoordinate atoms are, in general, susceptible to external perturbations. However, the present title clusters maintain the planarity and the attractive nature of the bonds even after multiple ligand binding at M centers in SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+. This would open the door for large-scale synthesis of phSi as well.Two major computational efforts were made before reaching our title phSi clusters. The first one is to examine SiE3M3+ (E = S–Po; M = Li–Cs) clusters, which adopt D3h or C3v structures as true minima (see Table S1 in ESI), being isoelectronic to the previous phC CE3M3+ (E = S–Po; M = Li–Cs) clusters. In the SiE3M3+ (E = S–Po; M = Li–Cs) clusters, the Si center always carries a positive charge ranging from 0.01 to +1.03|e|, in contrast to the corresponding phC species (see Fig. 1). Thus, electrostatic interactions between the Siδ+ and Mδ+ centers would be repulsive (Fig. 1). Given that the possibility of covalent interaction with an alkali metal is minimal, it would be a matter of debate whether they could be called true coordination. A second effort is to tune the electronegativity difference between Si and M centers so that the covalent contribution in Si–M bonding becomes substantial. Along this line, we consider the combinations of SiE3M3+ (E = N, P, As, Sb; M = Be, Mg, Ca, Sr, Ba). The results in Fig. S1 show that for E = Be and Mg, the phSi geometry has a large out-of-plane imaginary frequency mode, which indicates a size mismatch between the Si center and peripheral E3M3 (E = N–Bi; M = Be, Mg) ring. On the other hand, the use of larger M = Ca, Sr, Ba atoms effectively expands the size of the cavity and eventually leads to perfect planar geometry with Si atoms at the center as minima. In the case of SiBi3M3+, the planar isomer possesses a small imaginary frequency for M = Ca. Although planar SiBi3Sr3+ and SiBi3Ba3+ are true minima, they are 2.2 and 2.5 kcal mol−1 higher in energy than the lowest energy isomer, respectively (Fig. S2). Fig. 2 displays some selected low-lying isomers of SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba) clusters (see Fig. S3–S6 for additional isomers). The global minimum structure is a D3h symmetric phSi with an 1A1′ electronic state for all the twelve cases. The second lowest energy isomer, a ppSi, is located more than 49 kcal mol−1 above phSi for E = N. This relative energy between the most stable and nearest energy isomer gradually decreases upon moving from N to Sb. In the case of SiSb3M3+ clusters, the second-lowest energy isomer is 4.6–6.1 kcal mol−1 higher in energy than phSi. The nearest triplet state isomer is very high in energy (by 36–53 kcal mol−1, Fig. S3–S6) with respect to the global minimum.Open in a separate windowFig. 2The structures of low-lying isomers of SiE3M3+ (E = N, P, As, Sb; M = Ca, Sr, Ba) clusters. Relative energies (in kcal mol−1) are shown at the single-point CCSD(T)/def2-TZVP//PBE0/def2-TZVP level, followed by a zero-energy correction at PBE0. The values from left to right refer to Ca, Sr, and Ba in sequence. The group symmetries and electronic states are also given.Born–Oppenheimer molecular dynamics (BOMD) simulations at room temperature (298 K), taking SiE3Ca3+ clusters as case studies, were also performed. The results are displayed in Fig. S7. All trajectories show no isomerization or other structural alterations during the simulation time, as indicated by the small root mean square deviation (RMSD) values. The BOMD data suggest that the global minimum also has reasonable kinetic stability against isomerization and decomposition.The bond distances, natural atomic charges, and bond indices for SiE3Ca3+ clusters are given in for M = Sr, Ba). The Si–E bond distances are shorter than the typical Si–E single bond distance computed using the self-consistent covalent radii proposed by Pyykkö.27 In contrast, the Si–M bond distance is almost equal to the single bond distance. This gives the first hint of the presence of covalent bonding therein. However, the Wiberg bond indices (WBIs) for the Si–M links are surprisingly low (0.02–0.04). We then checked the Mayer bond order (MBO), which can be seen as a generalization of WBIs and is more acceptable since the approach of WBI calculations assumes orthonormal conditions of basis functions while the MBO considers an overlap matrix. The MBO values for the Si–M links are now sizable (0.13–0.18). These values are reasonable considering the large difference in electronegativity between Si and M, and, therefore, only a very polar bond is expected between them. In fact, the calculations of WBIs after orthogonalization of basis functions by the Löwdin method gives significantly large bond orders (0.48–0.55), which is known to overestimate the bond orders somewhat. The above results indicate that the presence of covalent bonding cannot be ruled out only by looking at WBI values.Bond distances (r, in Å), different bond orders (WBIs) {MBOs} [WBI in orthogonalized basis], and natural atomic charges (q, in |e|) of SiE3Ca3+ (E = N, P, As, Sb) clusters at the PBE0/def2-TZVP level
r Si–E r Si–Ca r E–Ca q Si q E q Ca
E = N1.6692.5552.2461.57−1.931.74
(1.14) {1.23} [1.84](0.02) {0.13} [0.51](0.22) {0.67} [0.84]
E = P2.1802.9352.6400.25−1.421.67
(1.34) {1.11} [1.52](0.03) {0.14} [0.54](0.27) {0.74} [1.05]
E = As2.3013.0042.7210.07−1.341.65
(1.33) {1.10} [1.45](0.03) {0.15} [0.55](0.29) {0.71} [1.12]
E = Sb2.5383.1552.896−0.39−1.161.62
(1.29) {1.01} [1.33](0.04) {0.18} [0.48](0.30) {0.78} [1.14]
Open in a separate windowOur following argument regarding the presence of covalent Si–M bonding is based on energy decomposition analysis (EDA) in combination with natural orbital for chemical valence (NOCV) theory. We first performed EDA by taking Ca and SiE3Ca2 in different charge and electronic states as interacting fragments to get the optimum fragmentation scheme that suits the best to describe the bonding situation (see Tables S6–S9). The size of orbital interaction (ΔEorb) is used as a probe.28 For all cases, Ca+ (D, 4s1) and SiE3Ca2 (D) in their doublet spin states turn out to be the best schemes, which give the lowest ΔEorb value.
Energy termInteractionCa+ (D, 4s1) + SiN3Ca2 (D)Ca+ (D, 4s1) + SiP3Ca2 (D)Ca+ (D, 4s1) + SiAs3Ca2 (D)Ca+ (D, 4s1) + SiSb3Ca2 (D)
ΔEint−192.9−153.0−144.9−129.9
ΔEPauli139.8115.2115.7110.9
ΔEelstata−162.0 (48.7%)−116.4 (43.4%)−113.0 (43.4%)−100.9 (41.9%)
ΔEorba−170.7 (51.3%)−151.8 (56.6%)−147.6 (56.6%)−140.0 (58.1%)
ΔEorb(1)bSiE3Ca2–Ca+(s) electron-sharing σ-bond−89.2 (52.3%)−79.4 (52.3%)−74.3 (50.3%)−66.9 (47.8%)
ΔEorb(2)bSiE3Ca2 → Ca+(d) π‖-donation−32.9 (19.3%)−32.0 (21.1%)−31.8 (21.5%)−30.8 (22.0%)
ΔEorb(3)bSiE3Ca2 → Ca+(d) σ-donation−13.1 (7.7%)−11.9 (7.8%)−12.0 (8.1%)−11.9 (8.5%)
ΔEorb(4)bSiE3Ca2 → Ca+(d) π-donation−12.3 (7.2%)−12.2 (8.0%)−12.5 (8.5%)−12.5 (8.9%)
ΔEorb(5)bSiE3Ca2 → Ca+(d) δ-donation−8.1 (4.7%)−9.9 (6.5%)−10.9 (7.4%)−11.8 (8.4%)
ΔEorb(rest)b−15.1 (8.8%)−6.4 (4.2%)−6.1 (4.1%)−6.1 (4.4%)
Open in a separate windowaThe values in parentheses are the percentage contributions to total attractive interactions (ΔEelstat + ΔEorb).bThe values in parentheses are the percentage contributions to the total orbital interaction ΔEorb.The decomposition of ΔEorb into pair-wise orbital interaction ΔEorb(n) in Fig. 3) helps us to identify the Si–Ca covalent bond and the orbitals involved in the pairwise interactions. The s orbital of Ca+ takes part in the electron-sharing σ-bond formation with SiE3Ca2, whereas vacant d AOs of Ca+ act as acceptor orbitals in the dative interactions, ΔEorb(2)–(5). Therefore, d AOs of Ca+ are responsible for 39–48% of the total orbital interaction. The present results further strengthen the proposal29–33 that heavier alkaline-earth elements (Ca, Sr, and Ba) should be classified as transition metals rather than main-group elements. Furthermore, a careful look at the Δρ(n) plots shows that in ΔEorb(1) and ΔEorb(2) only peripheral atoms are involved, but in ΔEorb(3)–(5) there is direct covalent interaction between Si and Ca centers. To correlate with the molecular orbitals (MOs) of the SiE3Ca3+ cluster, the related MOs for 24 valence electrons are given in Fig. S8. Δρ(3)–(5) can be correlated with HOMO-4, the HOMO and the HOMO′, respectively. Therefore, although the MO coefficient of Ca centers is small, they should not be neglected as the energy stabilization coming from them is significant. Si and M centers are only connected through delocalized bonds which is the reason for not having any gradient path between them as is indicated in the electron density analysis. Instead, there is a ring critical point at the center of the SiE2M ring (see Fig. S9). The results of adaptive natural density partitioning (AdNDP) analysis also corroborate this, where M centers are connected with the Si center through 7c–2e π-bonds (see Fig. S10).Open in a separate windowFig. 3Plot of the deformation densities, Δρ(1)–(5) corresponding to ΔEorb(1)–(5) and the related interacting orbitals of the fragments in the SiN3Ca3+ cluster at the PBE0/TZ2P-ZORA//PBE0/def2-TZVP level. The orbital energy values are in kcal mol−1. The charge flow of the deformation densities is from red to blue. The isovalue for Δρ(1) is 0.001 au and for the rest is 0.0005 au.Another aspect is to check the nature of electrostatic interaction between Si and M. The natural charges in ). Thus, the SiSb3M3+ cluster presents a case in which covalent bonding is robust and ionic interaction between Si and M centers is attractive in nature. If we look at the inter-atomic interaction energies (VTotal) for Si–M bonds and M–E bonds, it can be understood that the repulsive energy in Si–M bonds is largely overcompensated by two M–E bonds, even for E = N. This is the reason why electrostatic repulsion between Si and M centers does not result in a very large Si–M bond distance. Nevertheless, repulsive Si–M contacts in SiE3M3+ (E = N, P, As) make hexacoordination assignment skeptical. SiSb3M3+ clusters should be considered to possess phSi convincingly. Note that the IUPAC definition of coordination number only demands “the number of other atoms directly linked to that specified atom”,34 but does not say about the overall nature of interaction between them. In SiSb3M3+, phSi is linked to three Sb atoms through strong covalent bonds and is bound to three M atoms through ionic interaction in combination with a weaker covalent interaction. These clusters are only weakly aromatic because of such polar electronic distribution (see Fig. S11).The next challenge is to protect the reactive centers of phSi clusters with bulky ligands, which is required for large scale synthesis. This is not an easy task since slight external perturbation of most of the planar hypercoordinate atom species could result in a loss in planarity. Few years ago, the groups of Ding and Merino35 reported CAl4MX2 (M = Zr, Hf; X = F–I, C5H5) where ppC is sandwiched and protected by a metallocene framework. Therefore, the presence of X groups is mandatory to provide the electronic stabilization in ppC. In the present cases, surprisingly, SiSb3M3+ clusters are found to maintain the planarity around hexagons even after the coordination of M centers with six N-heterocyclic carbene (NHC) and benzene (Bz) ligands forming SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes, respectively (see Fig. 4). These complexes are highly stable against ligand dissociation as reflected by the high bond dissociation energy (De = 236.1 (Ca), 203.9 (Sr) and 171.3 (Ba) kcal mol−1) for SiSb3M3(NHC)6+ → SiSb3M3+ + 6NHC and De = 153.8 (Ca), 128.0 (Sr) and 114.0 (Ba) kcal mol−1 for SiSb3M3(Bz)6+ → SiSb3M3+ + 6Bz. The Si–M bond distances are slightly elongated because of coordination with the ligands. But the results of IQA given in Table S13 show that Si–M bonds have attractive interaction energies ranging between −20.0 and −32.4 kcal mol−1. Therefore, the planarity of the phSi core and the attractive nature of all the six contacts of phSi are maintained in ligand-bound SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes.Open in a separate windowFig. 4The minimum energy geometries of SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes at the PBE0-D3(BJ)/def2-TZVP level.In summary, we have theoretically achieved the first series of planar hexacoordinate silicon (phSi) clusters, SiSb3M3+ (M = Ca, Sr, Ba), by exploring their potential energy surfaces. These phSi systems are both thermodynamically and kinetically stable. The global minimum structures of SiE3M3+ (E = N, P, As, Sb) clusters have a D3h symmetry with the 1A1′ electronic state. The ability of the heavier alkaline-earth metals (Ca–Ba) to utilize their d orbitals in chemical bonding is a key factor that underlies the stability of these systems. The Ca–Ba ligands form weak covalent bonding with Si centers through their d orbitals, mimicking transition metals. The electronic charge distribution and IQA analysis show that electrostatic interaction in the Si–Ca links is essentially repulsive in SiN3M3+, but it sharply reduces with the decrease in electronegativity of E. Eventually, a sizable electrostatic attractive interaction exists between Si and M centers in SiSb3M3+, leading to a truly unprecedented phSi bonding motif that is held together by both covalent bonding and attractive ionic interaction. For SiE3M3+ (E = N, P, As) clusters, the electrostatic repulsion between Si and M dominates over covalent interaction, making Si–M contacts repulsive in nature. Most interestingly, the planarity of the phSi core and the attractive nature of all the six contacts of phSi are maintained in N-heterocyclic carbene (NHC) and benzene (Bz) bound SiSb3M3(NHC)6+ and SiSb3M3(Bz)6+ (M = Ca, Sr, Ba) complexes. Therefore, such clusters protected by bulky ligands would be suitable candidates for large scale synthesis in the presence of bulky counter-ions. Recent experimental reports on ptSi systems have already stimulated much curiosity within the community, and the present results would undoubtedly act as a stimulus to it.  相似文献   

7.
Deciphering the selectivity descriptors of heterogeneous metal phthalocyanine electrocatalysts for hydrogen peroxide production     
Yubo Yuan  Huan Li  Zhan Jiang  Zhichao Lin  Yirong Tang  Hongxuan Wang  Yongye Liang 《Chemical science》2022,13(37):11260
  相似文献   

8.
Axially chiral indolenine derived chromophore dimers and their chiroptical absorption and emission properties     
Emely Freytag  Marco Holzapfel  Asim Swain  Gerhard Bringmann  Matthias Stolte  Frank Würthner  Christoph Lambert 《Chemical science》2022,13(41):12229
Yamamoto homocoupling of two chiral oxindoles led to the atropo-diastereoselective formation of an axially chiral oxindole dimer. This building block served as the starting material for the syntheses of axially chiral squaraine and merocyanine chromophore dimers. These dimers show pronounced chiroptical properties, this is, outstandingly high ECD signals (Δε up to ca. 1500 M−1 cm−1) as a couplet with positive Cotton effect for the P-configuration around the biaryl axis and a negative Cotton effect for the M-configuration. All investigated dimers also exhibit pronounced circularly polarised emission with anisotropy values of ca. 10−3 cgs. Time-dependent density functional calculations were used to analyse the three contributions (local one electron, electric–magnetic coupling, and exciton coupling) to the rotational strength applying the Rosenfeld equation to excitonically coupled chromophores. While the exciton coupling term proves to be the dominant one, the electric–magnetic coupling possesses the same sign and adds significantly to the total rotational strength owing to a favourable geometric arrangement of the two chromophores within the dimer.

From an axially chiral oxindole, squaraine and merocyanine chromophore dimers with pronounced chiroptical properties were prepared.  相似文献   

9.
Temperature-dependence of radical-trapping activity of phenoxazine,phenothiazine and their aza-analogues clarifies the way forward for new antioxidant design     
Jia-Fei Poon  Luke A. Farmer  Evan A. Haidasz  Derek A. Pratt 《Chemical science》2021,12(33):11065
The prediction and/or rationalization of diarylamine radical-trapping antioxidant (RTA) activity at the elevated temperatures where they are most useful presents a significant challenge, precluding the development of new technology. Whilst structure–activity relationships at ambient temperatures are well established, their predictive capacity at elevated temperatures is poor due to competing reactions. A striking example involves phenoxazine, which is a superior RTA relative to its sulfur analog phenothiazine at ambient temperature (e.g. k = 3.9 × 107vs. 7.6 × 106 M−1 s−1 at 37 °C, respectively), but is demonstrably inferior at elevated temperatures. Despite being inherently less oxidizable in electrochemical experiments and high-accuracy computations, phenoxazine is more rapidly consumed than phenothiazine in autoxidations at 160 °C – a result which we attribute to a lower reorganization energy barrier to oxidation. Given these observations, we surmised that incorporation of an electronegative N-atom into the phenoxazine ring system would decrease the driving force for oxidation and ‘rescue’ its activity. Indeed, this was found to be the case for nitrogen incorporation at any position, regardless of the impact on the inherent RTA activity. Analogous experiments were carried out on phenothiazines into which nitrogen atoms were incorporated, revealing little benefit at 160 °C. These results suggest that for highly reactive diarylamines (i.e. those with k > 106 M−1 s−1), further enhancements in reactivity do not materially improve their ability to inhibit hydrocarbon autoxidation at elevated temperatures. Instead, their stability to one-electron oxidation determines their efficacy.

Studies of potent diarylamine antioxidants illustrate how their high-temperature activity can be limited by direct reaction with O2, and how driving force and barrier to oxidation can be modulated for future antioxidant development.  相似文献   

10.
Synthesis and photochemical properties of porphyrin-azobenzene triad     
Takashi Yamamura  Tatsuo Arai 《Tetrahedron letters》2004,45(50):9219-9223
Porphyrin-azobenzene triad E-2, having ester spacers between the chromophores, was synthesized and its photochemical and thermal isomerization properties were investigated. Triad E-2 showed little electronic communication among the chromophores according to a comparison of the steady-state absorption and fluorescence spectra of E-2 and their model compounds. E-2 showed photochemical E-Z isomerization and subsequent thermal Z-to-E isomerization. The porphyrin chromophore in Z-2, obtained by photoisomerization, did not strongly affect the transient state of thermal Z-to-E isomerization of the azobenzene unit. The high E-isomer composition at the photostationary state indicates the occurrence of triplet energy transfer between porphyrin and azobenzene. E-2 forms a 1:1 complex with 4,4′-bipyridyl (Bipy). The stability constant for E-2 with Bipy was determined (log K = 4.20 mol−1 dm3) by iterative least-squares fitting to a 1:1 binding model.  相似文献   

11.
Modulating the ground state,stability and charge transport in OFETs of biradicaloid hexahydro-diindenopyrene derivatives and a proposed method to estimate the biradical character     
Tanguy Jousselin-Oba  Masashi Mamada  Atsushi Okazawa  Jrome Marrot  Takayuki Ishida  Chihaya Adachi  Abderrahim Yassar  Michel Frigoli 《Chemical science》2020,11(44):12194
Biradicaloid compounds with an open-shell ground state have been the subject of intense research in the past decade. Although diindenoacenes are one of the most developed families, only a few examples have been reported as active layers in organic field-effect transistors (OFETs) with a charge mobility of around 10−3 cm2 V−1 s−1 due to a steric disadvantage of the mesityl group to kinetically stabilize compounds. Herein, we disclose our efforts to improve the charge transport of the diindenoacene family based on hexahydro-diindenopyrene (HDIP) derivatives with different annelation modes for which the most reactive position has been functionalized with (triisopropylsilyl)ethynyl (TIPS) groups. All the HDIP derivatives show remarkably higher stability than that of TIPS-pentacene, enduring for 2 days to more than 30 days, which depends on the oxidation potential, the contribution of the singlet biradical form in the ground state and the annelation mode. The annelation mode affects not only the band gap and the biradical character (y0) but also the value of the singlet–triplet energy gap (ΔES–T) that does not follow the reverse trend of y0. A method based on comparison between experimental and theoretical bond lengths has been disclosed to estimate y0 and shows that y0 computed at the projected unrestricted Hartree–Fock (PUHF) level is the most relevant among those reported by all other methods. Thanks to their high stability, thin-film OFETs were successfully fabricated. Well balanced ambipolar transport was obtained in the order of 10−3 cm2 V−1 s−1 in the bottom-gate/top-contact configuration, and unipolar transport in the top-gate/bottom-contact configuration was obtained in the order of 10−1 cm2 V−1 s−1 which is the highest value obtained for biradical compounds with a diindenoacene skeleton.

Biradicaloid HDIP derivatives show that the ΔES–T gap does not follow the reverse trend of the biradical character but depends more on the delocalization of the radical centres at the outer rings.  相似文献   

12.
Oxygen atom transfer promoted nitrate to nitric oxide transformation: a step-wise reduction of nitrate → nitrite → nitric oxide     
Kulbir  Sandip Das  Tarali Devi  Mrigaraj Goswami  Mahesh Yenuganti  Prabhakar Bhardwaj  Somnath Ghosh  Subash Chandra Sahoo  Pankaj Kumar 《Chemical science》2021,12(31):10605
Nitrate reductases (NRs) are molybdoenzymes that reduce nitrate (NO3) to nitrite (NO2) in both mammals and plants. In mammals, the salival microbes take part in the generation of the NO2 from NO3, which further produces nitric oxide (NO) either in acid-induced NO2 reduction or in the presence of nitrite reductases (NiRs). Here, we report a new approach of VCl3 (V3+ ion source) induced step-wise reduction of NO3 in a CoII-nitrato complex, [(12-TMC)CoII(NO3)]+ (2,{CoII–NO3}), to a CoIII–nitrosyl complex, [(12-TMC)CoIII(NO)]2+ (4,{CoNO}8), bearing an N-tetramethylated cyclam (TMC) ligand. The VCl3 inspired reduction of NO3 to NO is believed to occur in two consecutive oxygen atom transfer (OAT) reactions, i.e., OAT-1 = NO3 → NO2 (r1) and OAT-2 = NO2 → NO (r2). In these OAT reactions, VCl3 functions as an O-atom abstracting species, and the reaction of 2 with VCl3 produces a CoIII-nitrosyl ({CoNO}8) with VV-Oxo ({VV Created by potrace 1.16, written by Peter Selinger 2001-2019 O}3+) species, via a proposed CoII-nitrito (3, {CoII–NO2}) intermediate species. Further, in a separate experiment, we explored the reaction of isolated complex 3 with VCl3, which showed the generation of 4 with VV-Oxo, validating our proposed reaction sequences of OAT reactions. We ensured and characterized 3 using VCl3 as a limiting reagent, as the second-order rate constant of OAT-2 (k2/) is found to be ∼1420 times faster than that of the OAT-1 (k2) reaction. Binding constant (Kb) calculations also support our proposition of NO3 to NO transformation in two successive OAT reactions, as Kb(CoII–NO2) is higher than Kb(CoII–NO3), hence the reaction moves in the forward direction (OAT-1). However, Kb(CoII–NO2) is comparable to Kb{CoNO}8, and therefore sequenced the second OAT reaction (OAT-2). Mechanistic investigations of these reactions using 15N-labeled-15NO3 and 15NO2 revealed that the N-atom in the {CoNO}8 is derived from NO3 ligand. This work highlights the first-ever report of VCl3 induced step-wise NO3 reduction (NRs activity) followed by the OAT induced NO2 reduction and then the generation of Co-nitrosyl species {CoNO}8.

Single metal-induced reduction of NO3 → {NO2} → NO via oxygen atom transfer reaction.  相似文献   

13.
Phase-engineered high-entropy metastable FCC Cu2−yAgy(InxSn1−x)Se2S nanomaterials with high thermoelectric performance     
Wanjia Zhang  Yue Lou  Hongliang Dong  Fanshi Wu  Janak Tiwari  Zhan Shi  Tianli Feng  Sokrates T. Pantelides  Biao Xu 《Chemical science》2022,13(35):10461
Crystal-phase engineering to create metastable polymorphs is an effective and powerful way to modulate the physicochemical properties and functions of semiconductor materials, but it has been rarely explored in thermoelectrics due to concerns over thermal stability. Herein, we develop a combined colloidal synthesis and sintering route to prepare nanostructured solids through ligand retention. Nano-scale control over the unconventional cubic-phase is realized in a high-entropy Cu2−yAgy(InxSn1−x)Se2S (x = 0–0.25, y = 0, 0.07, 0.13) system by surface-ligand protection and size-driven phase stabilization. Different from the common monoclinic phase, the unconventional cubic-phase samples can optimize electrical and thermal properties through phase and entropy design. A high power factor (0.44 mW m−1 K−2), an ultralow thermal conductivity (0.25 W m−1 K−1) and a ZT value of 1.52 are achieved at 873 K for the cubic Cu1.87Ag0.13(In0.06Sn0.94)Se2S nanostructured sample. This study highlights a new method for the synthesis of metastable phase high-entropy materials and gives insights into stabilizing the metastable phase through ligand retention in other research communities.

The retention in size caused by the residual ligands drives the stability of metastable phase, enhancing structure symmetry and leading to good electrical transport. The distorted lattice and multidimensional defects intensify phonon scattering.  相似文献   

14.
Chromophore-radical excited state antiferromagnetic exchange controls the sign of photoinduced ground state spin polarization     
Martin L. Kirk  David A. Shultz  Patrick Hewitt  Daniel E. Stasiw  Ju Chen  Art van der Est 《Chemical science》2021,12(41):13704
A change in the sign of the ground-state electron spin polarization (ESP) is reported in complexes where an organic radical (nitronylnitroxide, NN) is covalently attached to a donor–acceptor chromophore via two different meta-phenylene bridges in (bpy)Pt(CAT-m-Ph-NN) (mPh-Pt) and (bpy)Pt(CAT-6-Me-m-Ph-NN) (6-Me-mPh-Pt) (bpy = 5,5′-di-tert-butyl-2,2′-bipyridine, CAT = 3-tert-butylcatecholate, m-Ph = meta-phenylene). These molecules represent a new class of chromophores that can be photoexcited with visible light to produce an initial exchange-coupled, 3-spin (bpy˙, CAT+˙ = semiquinone (SQ), and NN), charge-separated doublet 2S1 (S = chromophore excited spin singlet configuration) excited state. Following excitation, the 2S1 state rapidly decays to the ground state by magnetic exchange-mediated enhanced internal conversion via the 2T1 (T = chromophore excited spin triplet configuration) state. This process generates emissive ground state ESP in 6-Me-mPh-Pt while for mPh-Pt the ESP is absorptive. It is proposed that the emissive polarization in 6-Me-mPh-Pt results from zero-field splitting induced transitions between the chromophoric 2T1 and 4T1 states, whereas predominant spin–orbit induced transitions between 2T1 and low-energy NN-based states give rise to the absorptive polarization observed for mPh-Pt. The difference in the sign of the ESP for these molecules is consistent with a smaller excited state 2T14T1 gap for 6-Me-mPh-Pt that derives from steric interactions with the 6-methyl group. These steric interactions reduce the excited state pairwise SQ-NN exchange coupling compared to that in mPh-Pt.

A change in the sign of the ground state electron spin polarization (ESP) is reported in complexes where an organic radical (nitronylnitroxide, NN) is covalently attached to a donor–acceptor chromophore via two different meta-phenylene bridges.  相似文献   

15.
A small-molecule organic ferroelectric with piezoelectric voltage coefficient larger than that of lead zirconate titanate and polyvinylidene difluoride     
Han-Yue Zhang 《Chemical science》2022,13(17):5006
Piezoelectric materials that generate electricity when deforming are ideal for many implantable medical sensing devices. In modern piezoelectric materials, inorganic ceramics and polymers are two important branches, represented by lead zirconate titanate (PZT) and polyvinylidene difluoride (PVDF). However, PVDF is a nondegradable plastic with poor crystallinity and a large coercive field, and PZT suffers from high sintering temperature and toxic heavy element. Here, we successfully design a metal-free small-molecule ferroelectric, 3,3-difluorocyclobutanammonium hydrochloride ((3,3-DFCBA)Cl), which has high piezoelectric voltage coefficients g33 (437.2 × 10−3 V m N−1) and g31 (586.2 × 10−3 V m N−1), a large electrostriction coefficient Q33 (about 4.29 m4 C−2) and low acoustic impedance z0 (2.25 × 106 kg s−1 m−2), significantly outperforming PZT (g33 = 34 × 10−3 V m N−1 and z0 = 2.54 × 107 kg s−1 m−2) and PVDF (g33 = 286.7 × 10−3 V m N−1, g31 = 185.9 × 10−3 V m N−1, Q33 = 1.3 m4 C−2, and z0 = 3.69 × 106 kg s−1 m−2). Such a low acoustic impedance matches that of the body (1.38–1.99 × 106 kg s−1 m−2) reasonably well, making it attractive as next-generation biocompatible piezoelectric devices for health monitoring and “disposable” invasive medical ultrasound imaging.

A small-molecule organic ferroelectric (3,3-DFCBA)Cl has high piezoelectric voltage coefficients g33 (437.2 × 10−3 V m N−1), a large electrostriction coefficient Q33, and low acoustic impedance z0, far beyond that of PZT and PVDF.  相似文献   

16.
Plasmonic O2 dissociation and spillover expedite selective oxidation of primary C–H bonds     
Hao Li  Huan Shang  Fuze Jiang  Xingzhong Zhu  Qifeng Ruan  Lizhi Zhang  Jing Wang 《Chemical science》2021,12(46):15308
Manipulating O2 activation via nanosynthetic chemistry is critical in many oxidation reactions central to environmental remediation and chemical synthesis. Based on a carefully designed plasmonic Ru/TiO2−x catalyst, we first report a room-temperature O2 dissociation and spillover mechanism that expedites the “dream reaction” of selective primary C–H bond activation. Under visible light, surface plasmons excited in the negatively charged Ru nanoparticles decay into hot electrons, triggering spontaneous O2 dissociation to reactive atomic ˙O. Acceptor-like oxygen vacancies confined at the Ru–TiO2 interface free Ru from oxygen-poisoning by kinetically boosting the spillover of ˙O from Ru to TiO2. Evidenced by an exclusive isotopic O-transfer from 18O2 to oxygenated products, ˙O displays a synergistic action with native ˙O2 on TiO2 that oxidizes toluene and related alkyl aromatics to aromatic acids with extremely high selectivity. We believe the intelligent catalyst design for desirable O2 activation will contribute viable routes for synthesizing industrially important organic compounds.

Room-temperature O2 dissociation and spillover, as driven by plasmonic Ru on oxygen-deficient TiO2, expedite the selective oxidation of primary C–H bonds in alkyl aromatics for synthesizing industrially important organic compounds.  相似文献   

17.
Electro-optic crosslinkable chromophores with ultrahigh electro-optic coefficients and long-term stability     
Ziying Zeng  Jianhua Liu  Tongyu Luo  Zhibei Li  Juanfei Liao  Weijun Zhang  Lian Zhang  Fenggang Liu 《Chemical science》2022,13(45):13393
The development of organic electro-optic materials with ultrahigh electro-optic coefficients and high long-term alignment stability is the most challenging topic in this field. Next-generation crosslinkable nonlinear optical chromophore molecular glasses were developed to address this problem. A highly stable EO system including crosslinkable binary chromophores QLD1 and QLD2 or crosslinkable single chromophore QLD3 and multichromophore QLD4 with large hyperpolarizability was synthesized using tetrahydroquinoline as the donor. When the temperature continues to rise after poling, the chromophores modified with anthracene and acrylate can undergo Diels–Alder crosslinking reaction to fix the oriented chromophores through chemical bonds. After crosslinking, the QLD1/QLD2 and QLD2/QLD4 films achieved very high maximum r33 values of 327 and 373 pm V−1, respectively, which are the highest values reported for crosslinkable chromophore systems. After Diels–Alder cycloaddition, the glass transition temperature of the EO film increased by ∼90 °C to 185 °C, which is higher than for any other pure chromophore films. After being annealed at 85 °C, 99.63% of the initial r33 value could be maintained for over 500 h. The ultrahigh electro-optic activity and high long-term alignment stability of these materials showed new breakthroughs in organic EO materials for practical device explorations.

Ultrahigh electro-optic activity and high long-term alignment stability were achieved with crosslinkable binary chromophores QLD1 and QLD2 or crosslinkable single chromophore QLD3 and multichromophore QLD4 with large hyperpolarizability using tetrahydroquinoline as the donor.  相似文献   

18.
Regulating the assembly and expansion of the silver cluster from the Ag37 to Ag46 nanowheel driven by heteroanions     
Rakesh Kumar Gupta  Li Li  Zhi Wang  Bao-Liang Han  Lei Feng  Zhi-Yong Gao  Chen-Ho Tung  Di Sun 《Chemical science》2023,14(5):1138
Precise control over the shape and size of metal nanoclusters through anion template-driven self-assembly is one of the key scientific goals in the nanocluster community, however, it is still not understood comprehensively. In this work, we report the controllable synthesis and atomically precise structures of silver nanowheels Ag37 and Ag46, using homo (Cl ions) and heteroanion (Cl and CrO42− ions) template strategies, along with macrocyclic p-phenyl-thiacalix[4]arene and small iPrS ligands. Structural analyses revealed that in Ag37, Cl ions serve as both local and global templates, whereas CrO42− ions function as local and Cl ions as global templates in Ag46, resulting in a pentagonal nanowheel (Ag37) and a hexagonal (Ag46) nanowheel. The larger ionic size and more negative charges of CrO42− ions than Cl ions offer more coordination sites for the silver atoms and are believed to be the key factors that drive the nanowheel core to expand significantly. Also, by taking advantage of the deep cavity of thiacalix[4]arene with an extended phenyl group, Ag46 has been used as a host material for dye adsorption depending on the charge and size of organic dyes. The successful use of heteroanions to control the expansion of well-defined silver nanowheels fills the knowledge gap in understanding the directing role of heteroanions in dictating the shape and size of nanoclusters at the atomic level.

A heteroanion self-assembly strategy for regulating the shape and expanding the size of the silver cluster from Ag37 to Ag46 with a macrocyclic (p-phenyl-thiacalix[4]arene) ligand has been presented.  相似文献   

19.
Photoelectrochemical water oxidation improved by pyridine N-oxide as a mimic of tyrosine-Z in photosystem II     
Yong Zhu  Guoquan Liu  Ran Zhao  Hua Gao  Xiaona Li  Licheng Sun  Fei Li 《Chemical science》2022,13(17):4955
Artificial photosynthesis provides a way to store solar energy in chemical bonds with water oxidation as a major challenge for creating highly efficient and robust photoanodes that mimic photosystem II. We report here an easily available pyridine N-oxide (PNO) derivative as an efficient electron transfer relay between an organic light absorber and molecular water oxidation catalyst on a nanoparticle TiO2 photoanode. Spectroscopic and kinetic studies revealed that the PNO/PNO+˙ couple closely mimics the redox behavior of the tyrosine/tyrosyl radical pair in PSII in improving light-driven charge separation via multi-step electron transfer. The integrated photoanode exhibited a 1 sun current density of 3 mA cm−2 in the presence of Na2SO3 and a highly stable photocurrent density of >0.5 mA cm−2 at 0.4 V vs. NHE over a period of 1 h for water oxidation at pH 7. The performance shown here is superior to those of previously reported organic dye-based photoanodes in terms of photocurrent and stability.

Stable and high photocurrent for water oxidation was achieved by an organic dye-sensitized photoanode with a pyridine N-oxide derivative as an efficient electron relay between the chromophore and molecular water oxidation catalyst.  相似文献   

20.
Illuminating anti-hydrozirconation: controlled geometric isomerization of an organometallic species     
Theresa Hostmann  Tom&#x; Neveselý  Ryan Gilmour 《Chemical science》2021,12(31):10643
A general strategy to enable the formal anti-hydrozirconation of arylacetylenes is reported that merges cis-hydrometallation using the Schwartz Reagent (Cp2ZrHCl) with a subsequent light-mediated geometric isomerization at λ = 400 nm. Mechanistic delineation of the contra-thermodynamic isomerization step indicates that a minor reaction product functions as an efficient in situ generated photocatalyst. Coupling of the E-vinyl zirconium species with an alkyne unit generates a conjugated diene: this has been leveraged as a selective energy transfer catalyst to enable EZ isomerization of an organometallic species. Through an Umpolung metal–halogen exchange process (Cl, Br, I), synthetically useful vinyl halides can be generated (up to Z : E = 90 : 10). This enabling platform provides a strategy to access nucleophilic and electrophilic alkene fragments in both geometric forms from simple arylacetylenes.

A general strategy to enable the formal anti-hydrozirconation of arylacetylenes is reported that merges cis-hydrometallation using the Schwartz Reagent (Cp2ZrHCl) with a subsequent light-mediated geometric isomerization at λ = 400 nm.

The venerable Schwartz reagent (Cp2ZrHCl) is totemic in the field of hydrometallation,1 where reactivity is dominated by syn-selective M–H addition across the π-bond.2,3 This mechanistic foundation can be leveraged to generate well-defined organometallic coupling partners that are amenable to stereospecific functionalization. Utilizing terminal alkynes as readily available precursors,4 hydrozirconation constitutes a powerful strategy to generate E-configured vinyl nucleophiles that, through metal–halogen exchange, can be converted to vinyl electrophiles in a formal Umpolung process.5 Whilst this provides a versatile platform to access the electronic antipodes of the E-isomer, the mechanistic course of addition renders access to the corresponding Z-isomer conspicuously challenging. To reconcile the synthetic importance of this transformation with the intrinsic challenges associated with anti-hydrometallation and metallometallation,6 it was envisaged that a platform to facilitate geometric isomerization7 would be of value. Moreover, coupling this to a metal–halogen exchange would provide a simple Umpolung matrix to access both stereo-isomers from a common alkyne precursor (Fig. 1).Open in a separate windowFig. 1The stereochemical course of alkyne hydrometallation using the Schwartz reagent and an Umpolung platform to generate both stereo-isomers from a common alkyne precursor.Confidence in this conceptual blueprint stemmed from a report by Erker and co-workers, in which irradiating the vinyl zirconium species derived from phenyl acetylene (0.5 M in benzene) with a mercury lamp (Philips HPK 125 and Pyrex filter) induced geometric isomerization.8 Whilst Hg lamps present challenges in terms of safety, temperature regulation, cost and wavelength specificity, advances in LED technology mitigate all of these points. Therefore, a process of reaction development was initiated to generalize the anti-hydrozirconation of arylacetylenes. Crucial to the success of this venture was identifying the light-based activation mode that facilitates alkene isomerization. Specifically, it was necessary to determine whether this process was enabled by direct irradiation of the vinyl zirconium species, or if the EZ directionality results from a subsequent selective energy transfer process involving a facilitator. Several accounts of the incipient vinyl zirconium species reacting with a second alkyne unit to generate a conjugated diene have been disclosed.9,10 It was therefore posited that the minor by-product diene may be a crucial determinant in driving this isomerization (Fig. 2).Open in a separate windowFig. 2A working hypothesis for the light-mediated anti-hydrozirconation via selective energy transfer catalysis.To advance this working hypothesis and generalize the formal anti-hydrozirconation process, the reaction of Cp2ZrHCl with 1-bromo-4-ethynylbenzene (A-1) in CH2Cl2 was investigated ( for full details). This generates a versatile electrophile for downstream synthetic applications. Gratifyingly, after only 15 minutes, a Z : E-composition of 50 : 50 was reached (entry 1) and, following treatment with NBS, the desired vinyl bromide (Z)-1 was obtained in 76% yield (isomeric mixture) over the two steps. Further increasing the irradiation by 15 minute increments (entries 2–4) revealed that the optimum reaction time for the isomerization is 45 minutes (74%, Z : E = 73 : 27, entry 3). Extending the reaction time to 60 minutes (entry 4, 54%) did not lead to an improvement in selectivity and this was further confirmed by irradiating the reaction mixture for 90 minutes (entry 5). In both cases, a notable drop in yield was observed and therefore the remainder of the study was performed using the conditions described in entry 3. Next, the influence of the irradiation wavelength on the isomerization process was examined (entries 6–11). From a starting wavelength of λ = 369 nm, which gave a Z : E-ratio of 27 : 73 (entry 6), a steady improvement was observed by increasing the wavelength to λ = 374 nm (Z : E = 44 : 56, entry 7) and λ = 383 nm (Z : E = 53 : 47, entry 8). The selectivity reached a plateau at λ = 400 nm, with higher wavelengths proving to be detrimental (Z : E = 60 : 40 at λ = 414 nm, entry 9; Z : E = 26 : 74 at λ = 435 nm, entry 10). It is interesting to note that at λ = 520 nm, Z-1 was not detected by 1H NMR (entry 11).Reaction optimizationa
Entryλ [nm]Time [min]YieldbZ : E ratiob
14001576%50 : 50
24003072%68 : 32
34004574% (74%)74 : 26 (73 : 27)
44006054%73 : 27
54009049%73 : 27
63694566%27 : 73
73744561%44 : 56
83834564%53 : 47
94144567%60 : 40
104354572%26 : 74
115204567%<5 : 95
Open in a separate windowa(i) Cp2ZrHCl (62 mg, 0.24 mmol, 1.2 eq.), CH2Cl2 (1.5 mL), alkyne A-1 (36 mg, 0.2 mmol, 1.0 eq.) in CH2Cl2 (0.5 mL); (ii) irradiation; (iii) NBS (39 mg, 0.22 mmol, 1.1 eq.).baverage yield and Z : E ratio of two reactions determined by 1H-NMR with DMF as internal standard; isolated yield of the Z : E-mixture and Z : E-ratio in parentheses.Having identified standard conditions to enable a hydrozircononation/isomerization/bromination sequence, the scope and limitations of the method was explored using a range of electronically and structurally diverse phenylacetylenes (Fig. 3). This constitutes a net anti-Markovnikov hydrobromination of alkynes.11Open in a separate windowFig. 3Aromatic scope for the formal anti-hydrozirconation of terminal alkynes; reaction conditions: (i) Cp2ZrHCl (62 mg, 0.24 mmol, 1.2 eq.), CH2Cl2 (1.5 mL), alkyne A-1-17 (0.2 mmol, 1.0 eq.) in CH2Cl2 (0.5 mL), 15 min; (ii) irradiation (λ = 400 nm), 45 min; (iii) NBS (39 mg, 0.22 mmol, 1.1 eq.), 15 min; aisolated yield of Z : E-mixture as average of two reactions; b(i) Cp2ZrHCl (62 mg, 0.24 mmol, 1.2 eq.), CH2Cl2 (1.5 mL), alkyne A-15 (26 mg, 0.2 mmol, 1.0 eq.) in CH2Cl2 (0.5 mL); (ii) irradiation (λ = 400 nm), 45 min; (iii) PdPPh3 (7 mg, 0.006 mmol, 0.03 eq.) in THF (0.4 mL), BnBr (24 μL, 0.2 mmol, 1.0 eq.), rt, 18 h.12The introduction of halogen substituents in the 4-position proved to be compatible with the reaction conditions, enabling the formation of (Z)-1-4 in up to 81% yield (up to Z : E = 74 : 26). Interestingly, the introduction of the o-F (Z)-5 substituent led to a drop in the yield and selectivity: this is in stark contrast to cinnamoyl derivatives that have previously been examined in this laboratory.12 The m-Br proved to be less challenging enabling (Z)-6 to be generated smoothly (74%, Z : E = 67 : 33). The parent phenylacetylene (A-7) could be converted with a similar Z : E-ratio to (Z)-7 albeit less efficiently (36%, Z : E = 72 : 28). Electron donating groups in the para position such as (Z)-8-10 led to a general improvement in selectivity (up to 80%, Z : E = 81 : 19). Whereas methylation at the ortho-position compromised efficiency [(Z)-11, 37%, Z : E = 68 : 32], translocation to the meta-position led to a recovery in terms of yield and Z : E-ratio [(Z)-12, 71%, Z : E = 75 : 25]. Extending the π-system from phenyl to naphthyl enabled the generation of (Z)-13 90% and with a Z : E-ratio of 77 : 23. To enable a direct comparison of strongly and weakly donating groups on the reaction outcome the p-CF3 and p-OMe derivatives were examined. In the trifluoromethyl derivative (Z)-14 a decrease in yield (31%) and selectivity (Z : E = 48 : 52) was noted. In contrast, the para methoxy group in (Z)-15 led to an enhanced Z : E ratio of 86 : 14 (68% yield). This behavior was also observed with the trimethoxy derivative (Z)-16 (Z : E-ratio of 81 : 19). The piperonyl derivative performing similarly to the para methoxy derivative thereby enabling the formation of (Z)-17 with a Z : E-ratio of 85 : 15 (67% yield). Finally, to demonstrate the utility of the method, a direct transmetallation protocol was performed to intercept the Z-vinyl zirconium species with benzyl bromide.13 This enabled the synthesis of (Z)-18 in 67% yield.To demonstrate the compatibility of this platform with other common electrophiles, the deuterated, chlorinated and iodinated systems (Z)-19, -20 and -21 were prepared (Fig. 4). Yields and selectivities that are fully comparable with Fig. 3 were observed (up to 80% yield and Z : E = 80 : 20). Finally, to augment the photostationary composition further, a process of structural editing was conducted. It was envisaged that integrating a stabilizing non-covalent interaction in the Z-vinyl zirconium species may bias isomerization selectivity. Recent studies from this laboratory have established that a stabilizing interaction between the boron p-orbital and an adjacent non-bonding electron pair can be leveraged to induce a highly selective geometric isomerization of β-borylacrylates (Fig. 5, top).14Open in a separate windowFig. 4Scope of electrophiles for the formal anti-hydrozirconation; reaction conditions: (i) Cp2ZrHCl (62 mg, 0.24 mmol, 1.2 eq.), CH2Cl2 (1.5 mL), A-9 (36 mg, 0.2 mmol, 1.0 eq.) in CH2Cl2 (0.5 mL); (ii) irradiation (λ = 400 nm), 45 min; (iii) E+ (DCl, NCS or NIS) (0.22 mmol, 1.1 eq.), 15 min; isolated yields of the Z : E-mixture are reported.Open in a separate windowFig. 5Enhancing the selectivity of anti-hydrozirconation by leveraging a postulated nS → Zr interaction. Reaction conditions: (i) Cp2ZrHCl (62 mg, 0.24 mmol, 1.2 eq.), CH2Cl2 (1.5 mL), alkyne A-22-24 (0.2 mmol, 1.0 eq.) in CH2Cl2 (0.5 mL), rt, 15 min; (ii) irradiation (λ = 400 nm), 45 min; (iii) NBS (39 mg, 0.22 mmol, 1.1 eq.), rt, 15 min.Gratifyingly, the 5-bromo thiophenyl derivative (Z)-22 was generated with a Z : E ratio of 87 : 13 in 73% yield, and the unsubstituted derivative (Z)-23 was obtained in 41% yield higher selectivity (Z : E = 90 : 10). As a control experiment, the regioisomeric product (Z)-24 was prepared in which the sulfur atom is distal from the zirconium center. This minor alteration resulted in a conspicuous drop of selectivity (Z : E = 78 : 22), which is in line with the phenyl derivatives. Given the prominence of Frustrated-Lewis-Pairs (FLPs) in small molecule activation,15 materials such as (Z)-22 and (Z)-23 may provide a convenient starting point for the development of future candidates.To provide structural support for the formation of a Z-vinyl zirconium species upon irradiation at λ = 400 nm, the standard experiment was repeated in deuterated dichloromethane and investigated by 1H NMR spectroscopy. The spectra shown in Fig. 6 confirm the formation of transient E- and Z-vinyl zirconium species (E)-Zr1 and (Z)-Zr1 and are in good agreement with literature values.8 Diagnostic resonances of (E)-Zr1 include H1 at 7.76 ppm, whereas the analogous signal in (Z)-Zr1 is high field shifted to 6.33 ppm (Δδ(H1Z−E) = −1.43 ppm). In contrast, the H2 signal for (Z)-Zr1 appears at 7.56 ppm, which is at lower field compared to the H2 signal for (E)-Zr1 at 6.64 ppm (Δδ(H2Z−E) = 0.92 ppm). In the 13C-NMR spectra (see the ESI) the carbon signal of C1 and C2 are both low field shifted for (Z)-Zr1 compared to (E)-Zr1 (Δδ(C1Z−E) = 10.5 ppm and Δδ(C1Z-E) = 5.6 ppm).Open in a separate windowFig. 61H-NMR of the transient vinylzirconium species (E)-Zr1 (top) and (Z)-Zr1 (bottom).A computational analysis of the vinyl zirconium isomers (E)-Zr1 and (Z)-Zr1 revealed two low energy conformers for each geometry (Fig. 7. For full details see the ESI). These optimized structures served as a basis for more detailed excited state calculations using a time-dependent density functional theory (TDDFT) approach. These data indicate that isomerization of the styrenyl zirconium species by direct irradiation is highly improbable using λ = 400 nm LEDs. However, upon measuring the absorption spectrum of the reaction mixture (Fig. 8, bottom), the shoulder of a band reaching to the visible part of the spectrum is evident (for more details see the ESI). Furthermore, the fluorescence spectrum (Fig. 8, top) clearly shows light emission from the reaction mixture. Collectively, these data reinforce the working hypothesis that a minor reaction product functions as a productive sensitizer, thereby enabling the isomerization to occur via selective energy transfer.Open in a separate windowFig. 7A comparative analysis of (E)-Zr1 and (Z)-Zr1.Open in a separate windowFig. 8(Top) Fluorescence spectra of the reaction mixture before and after irradiation, and the diene 25 (c = 0.1 mm, irradiation at λ = 350 nm). (Bottom) Absorption spectra of the reaction mixture before and after irradiation (c = 0.1 mm), the alkyne A-1 and the diene 25 (c = 0.05 mm).As previously highlighted, phenylacetylenes are known to dimerize in the presence of Cp2Zr* based complexes.9,16 Therefore, to provide support for the involvement of such species, diene 25 was independently prepared and its absorption and emission spectra were compared with those of the reaction mixture (Fig. 8). The emission spectra of the reaction mixture and of diene 25 are closely similar. It is also pertinent to note that diene 25 was also detected in the crude reaction mixture by HRMS (see the ESI).Whilst the spectral measurements in Fig. 8 are in line with diene 25 functioning as an in situ photocatalyst, more direct support was desirable. Frustratingly, efforts to subject (E)-Zr-1 and (Z)-Zr-1 to standard Stern–Volmer quenching studies were complicated by difficulties in removing diene 25 from the samples. It was therefore envisaged that doping reactions with increasing quantities of diene 25 might be insightful. To that end, the hydrozirconation/isomerization sequence was performed with 0.5, 1.0 and 2.5 mol% of diene 25 and the reactions were shielded from light after 5 minutes. Analysis of the mixture by 1H NMR spectroscopy revealed a positive impact of 25 on the Z : E selectivity, (Z : E = 23 : 77, 24 : 76 and 30 : 70, respectively. Fig. 9, top). To further demonstrate the ability of diene 25 to act as an energy transfer catalyst for geometric isomerization, two model alkenes containing the styrenyl chromophore were exposed to the standard reaction conditions and the photostationary composition was measured after 45 min. Exposing trans-stilbene (E)-26 to the isomerization conditions furnished a Z : E photostationary composition of 44 : 56. Similarly, trans-β-methyl styrene (E)-27 could be isomerized to the cis-β-methyl styrene (Z)-27 with a Z : E ratio of 47 : 53. No isomerization was observed at λ = 400 nm in the absence of the catalyst. Whilst direct comparison with the isomerization of vinyl zirconium species must be made with caution, these experiments demonstrate that dienes such as 25 have the capacity to act as photosensitizers with styrenyl chromophores.Open in a separate windowFig. 9(Top) Exploring the impact of adding diene 25 as an external photocatalyst. (Bottom) Validating photosensitization of the styrenyl chromophore using diene 25.Collectively, these data support the hypothesis that isomerization does not result from direct irradiation alone,17 but that conjugated dienes, which are produced in small amounts, function as in situ energy transfer catalysts (Fig. 10). This antenna undergoes rapid inter-system crossing (ISC)18 to generate the triplet state and, upon energy transfer to the alkene fragment, returns to the ground state.19 This mechanistic study has guided the development of an operationally simple anti-hydrozirconation of alkynes that relies on inexpensive LED irradiation. Merging this protocol with a sequential metal–halogen exchange enables the formal anti-Markovnikov hydrobromination of alkynes11 and provides a sterodivergent platform to access defined alkene vectors from simple alkynes. This complements existing strategies to isomerize vinyl bromides,20 and circumvents the risks of vinyl cation formation and subsequent degradation.21 Finally, the selectivity of this geometric isomerization can be further augmented through the judicious introduction of stabilizing non-covalent interactions (up to Z : E = 90 : 10). It is envisaged that this selective, controlled geometric isomerization of an organometallic species will find application in contemporary synthesis. Furthermore, it contributes to a growing body of literature that describes the in situ formation of photoactive species upon irradiation.22Open in a separate windowFig. 10Postulated energy transfer catalysis cycle predicated on in situ formation of a conjugated diene photocatalyst.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号