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 93 毫秒
1.
Radical cation salts with halide mercurate anions were obtained by the electrochemical oxidation of bis(ethylenedioxo)tetrathiafulvalene (BEDO-TTF), and their conductivity was studied. The compounds (BEDO-TTF)4Hg3X8 (X = Cl or Br), (BEDO-TTF)4Hg3.5I9, and (BEDO-TTF)2HgBr3 possess the conductivity of the metallic type down to helium temperatures.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 5, pp. 1167–1170, May, 1996.  相似文献   

2.
Dong Shao  Xin‐Yi Wang 《中国化学》2020,38(9):1005-1018
Single‐molecule magnets (SMMs) are paramagnetic molecules that can be magnetized below a certain temperature and have potential applications in high‐density information storage, magnetic qubits, spintronic devices, etc. The discovery of the first SMM, Mn12, opened a new era of molecular magnetism and promoted collaborative researches between chemists and physicists for their exotic quantum as well as classical magnetic properties. In the recent past, great efforts have been made to develop strategies for constructing new SMMs with high energy barriers (Ueff) and blocking temperatures (TB), resulting in great and fast development of SMMs. In this concise review, we highlight the main synthetic approaches and representative results in the design and synthesis of high performance SMMs. We hope to give the readers a basic understanding of SMMs and a snapshot of the representative researches on SMMs from a perspective of synthetic chemists.  相似文献   

3.
High proton conducting electrolytes with mechanical moldability are a key material for energy devices. We propose an approach for creating a coordination polymer (CP) glass from a protic ionic liquid for a solid-state anhydrous proton conductor. A protic ionic liquid (dema)(H2PO4), with components which also act as bridging ligands, was applied to construct a CP glass (dema)0.35[Zn(H2PO4)2.35(H3PO4)0.65]. The structural analysis revealed that large Zn–H2PO4/H3PO4 coordination networks formed in the CP glass. The network formation results in enhancement of the properties of proton conductivity and viscoelasticity. High anhydrous proton conductivity (σ = 13.3 mS cm−1 at 120 °C) and a high transport number of the proton (0.94) were achieved by the coordination networks. A fuel cell with this CP glass membrane exhibits a high open-circuit voltage and power density (0.15 W cm−2) under dry conditions at 120 °C due to the conducting properties and mechanical properties of the CP glass.

A proton-conducting coordination polymer glass derived from a protic ionic liquid works as a moldable solid electrolyte and the anhydrous fuel cell showed IV performance of 0.15 W cm−2 at 120 °C.  相似文献   

4.
We present the magnetic properties of the metal-organic framework {[CoCxAPy]·2.15 H2O}n (Cx = bis(carboxypropyl)tetramethyldisiloxane; APy = 4,4′-azopyridine) (1) that builds up from the stacking of 2D coordination polymers. The 2D-coordination polymer in the bc plane is formed by the adjacent bonding of [CoCxAPy] 1D two-leg ladders with Co dimer rungs, running parallel to the c-axis. The crystal packing of 2D layers shows the presence of infinite channels running along the c crystallographic axis, which accommodate the disordered solvate molecules. The Co(II) is six-coordinated in a distorted octahedral geometry, where the equatorial plane is occupied by four carboxylate oxygen atoms. Two nitrogen atoms from APy ligands are coordinated in apical positions. The single-ion magnetic anisotropy has been determined by low temperature EPR and magnetization measurements on an isostructural compound {[Zn0.8Co0.2CxAPy]·1.5 CH3OH}n (2). The results show that the Co(II) ion has orthorhombic anisotropy with the hard-axis direction in the C2V main axis, lying the easy axis in the distorted octahedron equatorial plane, as predicted by the ab initio calculations of the g-tensor. Magnetic and heat capacity properties at very low temperatures are rationalized within a S* = 1/2 magnetic dimer model with anisotropic antiferromagnetic interaction. The magnetic dimer exhibits slow relaxation of the magnetization (SMM) below 6 K in applied field, with a τlf ≈ 2 s direct process at low frequencies, and an Orbach process at higher frequencies with U/kB = 6.7 ± 0.5 K. This compound represents a singular SMM MOF built-up of Co-dimers with an anisotropic exchange interaction.  相似文献   

5.
Dy-based single-molecule magnets (SMMs) are of great interest due to their ability to exhibit very large thermal barriers to relaxation and therefore high blocking temperatures. One interesting line of investigation is Dy-encapsulating endohedral clusterfullerenes, in which a carbon cage protects magnetic Dy3+ ions against decoherence by environmental noise and allows for the stabilization of bonding and magnetic interactions that would be difficult to achieve in other molecular architectures. Recent studies of such materials have focused on clusters with two Dy atoms, since ferromagnetic exchange between Dy atoms is known to reduce the rate of magnetic relaxation via quantum tunneling. Here, two new dysprosium-containing mixed-metallic sulfide clusterfullerenes, DyScS@Cs(6)–C82 and DyScS@C3v(8)–C82, have been successfully synthesized, isolated and characterized by mass spectrometry, Vis-NIR, cyclic voltammetry, single crystal X-ray diffractometry, and magnetic measurements. Crystallographic analyses show that the conformation of the encapsulated cluster inside the fullerene cages is notably different than in the Dy2X@Cs(6)–C82 and Dy2X@C3v(8)–C82 (X = S, O) analogues. Remarkably, both isomers of DyScS@C82 show open magnetic hysteresis and slow magnetic relaxation, even at zero field. Their magnetic blocking temperatures are around 7.3 K, which are among the highest values reported for clusterfullerene SMMs. The SMM properties of DyScS@C82 far outperform those of the dilanthanide analogues Dy2S@C82, in contrast to the trend observed for carbide and nitride Dy clusterfullerenes.

Extremely high magnetic blocking temperatures (∼7.3 K) were observed for DyScS endohedral fullerene single-molecule magnets.  相似文献   

6.
Hybrid compounds containing organic layers composed of tetrathiafulvalenes (BEDT-TTF, BETS, BEDO-TTF) and inorganic layers consisting of anions based on lead bromide were synthesized. The crystal structure of (BEDT-TTF)6Pb3Br10(PhBr) was determined. The temperature dependences of electrical conductivity for the synthesized compounds were measured and the EPR spectra of these compounds were recorded. In the structure of (BEDT-TTF)6Pb3Br10(PhBr), the organic conducting layers of BEDT-TTF molecules alternate with non-conducting layers composed of infinite chains of bromoplumbate anions [Pb3Br9]3–, solvent molecules, and Br anions. The newly synthesized BEDT-TTF bromoplumbates have similar linewidths of EPR signals, which indicate that their conducting layers have similar structures. The BEDT-TTF bromoplumbates are semiconductors, while temperature-dependent electrical resistivity measurements show the metallic behavior for BEDO-TTF and BETS bromoplumbates.  相似文献   

7.
The first three-dimensional (3D) conductive single-ion magnet (SIM), (TTF)2[Co(pdms)2] (TTF=tetrathiafulvalene and H2pdms=1,2-bis(methanesulfonamido)benzene), was electrochemically synthesised and investigated structurally, physically, and theoretically. The similar oxidation potentials of neutral TTF and the molecular precursor [HNEt3]2[M(pdms)2] (M=Co, Zn) allow for multiple charge transfers (CTs) between the SIM donor [M(pdms)2]n and the TTF.+ acceptor, as well as an intradonor CT from the pdms ligand to Co ion upon electrocrystallisation. Usually TTF functions as a donor, whereas in our system TTF is both a donor and an accepter because of the similar oxidation potentials. Furthermore, the [M(pdms)2]n donor and TTF.+ acceptor are not segregated but strongly interact with each other, contrary to reported layered donor–acceptor electrical conductors. The strong intermolecular and intramolecular interactions, combined with CT, allow for relatively high electrical conductivity even down to very low temperatures. Furthermore, SIM behaviour with slow magnetic relaxation and opening of hysteresis loops was observed. (TTF)2[Co(pdms)2] ( 2-Co ) is an excellent building block for preparing new conductive SIMs.  相似文献   

8.
Using the redox-active tetrathiafulvalene tetrabenzoate (TTFTB4−) as the linker, a series of stable and porous rare-earth metal–organic frameworks (RE-MOFs), [RE93-OH)133-O)(H2O)9(TTFTB)3] ( 1-RE , where RE=Y, Sm, Gd, Tb, Dy, Ho, and Er) were constructed. The RE93-OH)133-O) (H2O)9](CO2)12 clusters within 1-RE act as segregated single-molecule magnets (SMMs) displaying slow relaxation. Interestingly, upon oxidation by I2, the S=0 TTFTB4− linkers of 1-RE were converted into S= TTFTB.3− radical linkers which introduced exchange-coupling between SMMs and modulated the relaxation. Furthermore, the SMM property can be restored by reduction in N,N-dimethylformamide. These results highlight the advantage of MOFs in the construction of redox-switchable SMMs.  相似文献   

9.
The first three‐dimensional (3D) conductive single‐ion magnet (SIM), (TTF)2[Co(pdms)2] (TTF=tetrathiafulvalene and H2pdms=1,2‐bis(methanesulfonamido)benzene), was electrochemically synthesised and investigated structurally, physically, and theoretically. The similar oxidation potentials of neutral TTF and the molecular precursor [HNEt3]2[M(pdms)2] (M=Co, Zn) allow for multiple charge transfers (CTs) between the SIM donor [M(pdms)2]n? and the TTF.+ acceptor, as well as an intradonor CT from the pdms ligand to Co ion upon electrocrystallisation. Usually TTF functions as a donor, whereas in our system TTF is both a donor and an accepter because of the similar oxidation potentials. Furthermore, the [M(pdms)2]n? donor and TTF.+ acceptor are not segregated but strongly interact with each other, contrary to reported layered donor–acceptor electrical conductors. The strong intermolecular and intramolecular interactions, combined with CT, allow for relatively high electrical conductivity even down to very low temperatures. Furthermore, SIM behaviour with slow magnetic relaxation and opening of hysteresis loops was observed. (TTF)2[Co(pdms)2] ( 2‐Co ) is an excellent building block for preparing new conductive SIMs.  相似文献   

10.
The ring-opening oxidative amination of methylenecyclopropanes (MCPs) with diazenes catalyzed by py3TiCl2(NR) complexes is reported. This reaction selectively generates branched α-methylene imines as opposed to linear α,β-unsaturated imines, which are difficult to access via other methods. Products can be isolated as the imine or hydrolyzed to the corresponding ketone in good yields. Mechanistic investigation via density functional theory suggests that the regioselectivity of these products results from a Curtin–Hammett kinetic scenario, where reversible β-carbon elimination of a spirocyclic [2 + 2] azatitanacyclobutene intermediate is followed by selectivity-determining β-hydrogen elimination of the resulting metallacycle. Further functionalizations of these branched α-methylene imine products are explored, demonstrating their utility as building blocks.

The ring-opening oxidative amination of methylenecyclopropanes (MCPs) with diazenes catalyzed by py3TiCl2(NR) complexes is reported.  相似文献   

11.
How far can we push the limits in removing stereoelectronic protection from an unstable intermediate? We address this question by exploring the interplay between the primary and secondary stereoelectronic effects in the Baeyer–Villiger (BV) rearrangement by experimental and computational studies of γ-OR-substituted γ-peroxylactones, the previously elusive non-strained Criegee intermediates (CI). These new cyclic peroxides were synthesized by the peroxidation of γ-ketoesters followed by in situ cyclization using a BF3·Et2O/H2O2 system. Although the primary effect (alignment of the migrating C–Rm bond with the breaking O–O bond) is active in the 6-membered ring, weakening of the secondary effect (donation from the OR lone pair to the breaking C–Rm bond) provides sufficient kinetic stabilization to allow the formation and isolation of stable γ-hydroperoxy-γ-peroxylactones with a methyl-substituent in the C6-position. Furthermore, supplementary protection is also provided by reactant stabilization originating from two new stereoelectronic factors, both identified and quantified for the first time in the present work. First, an unexpected boat preference in the γ-hydroperoxy-γ-peroxylactones weakens the primary stereoelectronic effects and introduces a ∼2 kcal mol−1 Curtin–Hammett penalty for reacquiring the more reactive chair conformation. Second, activation of the secondary stereoelectronic effect in the TS comes with a ∼2–3 kcal mol−1 penalty for giving up the exo-anomeric stabilization in the 6-membered Criegee intermediate. Together, the three new stereoelectronic factors (inverse α-effect, misalignment of reacting bonds in the boat conformation, and the exo-anomeric effect) illustrate the richness of stereoelectronic patterns in peroxide chemistry and provide experimentally significant kinetic stabilization to this new class of bisperoxides. Furthermore, mild reduction of γ-hydroperoxy-γ-peroxylactone with Ph3P produced an isolable γ-hydroxy-γ-peroxylactone, the first example of a structurally unencumbered CI where neither the primary nor the secondary stereoelectronic effect are impeded. Although this compound is relatively unstable, it does not undergo the BV reaction and instead follows a new mode of reactivity for the CI – a ring-opening process.

Protecting stereoelectronic effects prevent Baeyer–Villiger rearrangement and stabilize γ-OX-γ-peroxylactones (X = H, OH), the previously elusive non-strained Criegee intermediates.  相似文献   

12.
Picosecond to millisecond laser time-resolved transient absorption spectroscopy was used to study molecular oxygen (O2) rebinding and conformational relaxation following O2 photodissociation in the α and β subunits within human hemoglobin in the quaternary R-like structure. Oxy-cyanomet valency hybrids, α2(Fe2+–O22(Fe3+–CN) and α2(Fe3+–CN)β2(Fe2+–O2), were used as models for oxygenated R-state hemoglobin. An extended kinetic model for geminate O2 rebinding in the ferrous hemoglobin subunits, ligand migration between the primary and secondary docking site(s), and nonexponential tertiary relaxation within the R quaternary structure, was introduced and discussed. Significant functional non-equivalence of the α and β subunits in both the geminate O2 rebinding and concomitant structural relaxation was revealed. For the β subunits, the rate constant for the geminate O2 rebinding to the unrelaxed tertiary structure and the tertiary transition rate were found to be greater than the corresponding values for the α subunits. The conformational relaxation following the O2 photodissociation in the α and β subunits was found to decrease the rate constant for the geminate O2 rebinding, this effect being more than one order of magnitude greater for the β subunits than for the α subunits. Evidence was provided for the modulation of the O2 rebinding to the individual α and β subunits within human hemoglobin in the R-state structure by the intrinsic heme reactivity through a change in proximal constraints upon the relaxation of the tertiary structure on a picosecond to microsecond time scale. Our results demonstrate that, for native R-state oxyhemoglobin, O2 rebinding properties and spectral changes following the O2 photodissociation can be adequately described as the sum of those for the α and β subunits within the valency hybrids. The isolated β chains (hemoglobin H) show similar behavior to the β subunits within the valency hybrids and can be used as a model for the β subunits within the R-state oxyhemoglobin. At the same time, the isolated α chains behave differently to the α subunits within the valency hybrids.

O2 rebinding and conformational relaxation following O2 photodissociation were studied on picosecond to millisecond time scale in the α and β subunits within human hemoglobin in the quaternary R-like structure.  相似文献   

13.
We report on the first isolation and structural characterization of an iron phosphinoimino-borane complex Cp*Fe(η2-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NC6H4PPh2) by dehydrogenation of iron amido-borane precursor Cp*Fe(η1-H3B–NHC6H4PPh2). Significantly, regeneration of the amido-borane complex has been realized by protonation of the iron(ii) imino-borane to the amino-borane intermediate [Cp*Fe(η2-H2B–NHC6H4PPh2)]+ followed by hydride transfer. These new iron species are efficient catalysts for 1,2-selective transfer hydrogenation of quinolines with ammonia borane.

Dehydrogenation of an amido-borane iron complex provides an imino-borane complex. Regeneration of the amido-borane precursor was achieved by protonation of the imino-borane followed by hydride transfer to the amino-borane intermediate.

Because of relevance to H2 storage1–10 and hydrogenation catalysis,11–15 metal amine-borane complexes16–18 and their dehydrogenated forms, such as amino-boranes20–22 and imino-boranes4 are arising as a significant family in organometallic chemistry. In transition metal-catalyzed dehydrocoupling of amine-boranes and related transfer hydrogenations, the interactions between the metal and the borane fragment are essential to dehydrogenation and the consequent transformations.16–20 Specifically, amino-borane complexes containing a M–H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2 moiety are the primary dehydrogenated species and are often identified as a resting point in the catalysis (Scheme 1a).20–22 Management of reversible dehydrogenation–regeneration reactions on a M–BH2 Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2 platform could provide a strategy with which to design efficient catalysts capable of operating sustainable syntheses.Open in a separate windowScheme 1Schematic representation of metal-based amine-borane dehydrogenation.Wider exploration of metal amino-borane chemistry is challenging since M–H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NH2 species are very reactive toward H2 release. In 2010, Aldridge et al. reported the isolation of [(IMes)2Rh(H)22-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2)] and [(IMes)2Ir(H)22-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2)] from the metal-catalyzed dehydrogenation of R2HN·BH3.21a At the same time, Alcaraz and Sabo-Etienne reported the preparation of (PCy3)2Ru(H)22-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NHnMe2−n) (n = 0–2) complexes22a by the dehydrogenation of amine-boranes with the corresponding ruthenium precursors. Subsequently, a straightforward synthesis of Ru, Rh, and Ir amino-borane complexes by reaction of H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2 (R = iPr or Cy) with the bis(hydrogen) complexes of M(H)22-H2)2(PCy3)2 or [CpRu(PR3)2]+ fragments was developed.21b,22b Turculet et al. have shown that the ruthenium-alkoxide complex is able to activate H3B·NHR2 producing hydrido ruthenium complex.23 Notably, Weller and Macgregor found that dehydrocoupling of ammonia-borane by [Ph2P(CH2)3PPh2Rh(η6-C6H5F)] affords a μ-amino-borane bimetallic Rh complex, in which the simplest H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NH2 moiety is trapped on a rhodium dimer.20aAlthough iron-catalyzed dehydrocoupling of amine-boranes has attracted great interest,24–29 iron amine-borane complexes, their dehydrogenated derivatives, and especially the catalysis relevant to organic synthesis are largely unexplored. Recently, Kirchner et al. reported a pincer-type iron complex generated by protonation of the borohydride iron complex (PNP)Fe(H)(η2-BH4) with ammonium salts.30 Inspired by earlier research on M–H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NR2 chemistry, we intended to establish the reversible conversions of amino-borane complexes and their dehydrogenated forms in a synthetic piano-stool iron system. Herein, we report dehydrogenation of iron amido-borane complex Cp*Fe(η1-H3B–NHC6H4PPh2) (2) (Cp* = Me5C5) to the imino-borane complex Cp*Fe(η2-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NC6H4PPh2) (3), and resaturation of the imino-borane by stepwise protonation and hydride transfer (Scheme 1b). This new class of iron species is capable of catalyzing 1,2-selective transfer hydrogenation of quinolines with H3N·BH3.To synthesize the iron amido-borane complex, a new monomer, the iron tetrahydridoborate precursor Cp*Fe(η2-BH4)(NCMe) (1), was prepared in situ by the reaction of [Cp*Fe(NCMe)3]PF6 with Bu4NBH4 in acetonitrile at room temperature for 5 min. Such ferrous borohydrides have been documented only rarely,31 since they are prone to form polynuclear iron borate clusters.32,33 The 11B NMR spectrum of the reaction solution shows a quintet at δ 15.4 (JBH = 88 Hz) for the BH4 ligand of 1, and this stands in contrast to the signal at δ −32.0 observed for Bu4NBH4. Upon storing the reaction mixture at −30 °C overnight, single crystals suitable for X-ray diffraction were obtained. Crystallographic analysis confirmed the structure of 1 as a piano-stool iron tetrahydridoborate compound (ESI, Fig. S1).Addition of phosphinoamine ligand 1,2-Ph2PC6H4NH2 to a solution of 1 in acetonitrile caused an instantaneous color change from deep blue to dark brown (Scheme 2). ESI-MS studies indicated the production of the iron amido-borane compound (2) with m/z = 481.1793 (calcd m/z = 481.1770), which was isolated in 87% yield. NMR spectra showed a boron resonance at δ −17.5, and a phosphorus resonance at δ 85.9. The 1H NMR spectrum exhibits a characteristic hydride signal at δ −13.98, which is assigned to the bridging hydride Fe–H–B. Owing to exchange between the hydrogen atoms at the boron,34 the terminal B–H resonances in the 1H NMR spectrum are very broad and are obscured by the distinct Cp* signals. To assign the B–H hydride signals, the deuterated compound Cp*Fe(D3B–NHC6H4PPh2) (d-2) was synthesized from Cp*Fe(BD4)(NCMe). In addition to the Fe–D–B signal at δ −13.98, the 2H NMR spectrum of d-2 displayed discrete peaks at δ 2.23 and 0.19 for the terminal B–D hydrides (Fig. 1).Open in a separate windowFig. 1 2H NMR spectra for dehydrogenation of d-2 to d-3.Open in a separate windowScheme 2Synthetic route to imino-borane complex.When a C6H6 solution of 2 was held at 50 °C for 6 h the dehydrogenated imino-borane compound (3) was produced in 92% yield. The ESI-MS spectrum of 3 has a strong peak at m/z 479.1626 (calcd m/z = 479.1637) which can be compared to the peak at m/z = 481.1793 for 2. The isotopic distributions match well with the calculated values (see Fig. S3). GC analysis shows that the reaction produced H2 nearly quantitatively (see Fig. S4). In solution, the 31P NMR spectrum of 3 displays a sharp signal at δ 71.9, in contrast to the peak at δ 85.9 for 2. The 11B resonance shifts significantly, from δ −17.5 for 2 to δ 42.7 for 3 (Fig. S16), and is particularly diagnostic of a three-coordinate boron atom.21,35 This result indicates the B Created by potrace 1.16, written by Peter Selinger 2001-2019 N double bond character in the dehydrogenated form of the amido-borane complex. In the 1H NMR spectrum, the Fe–H–B signal was observed at δ −17.91 with the integral of 2H, and no characteristic signal for a terminal B–H hydride was found. To confirm the formation of an imino-borane compound, the hydrogen decoupling was also carried out with compound d-2 and monitored by 2H NMR spectra. Only a deuterium signal was observed at δ −17.91 for Fe–D–B, indicating the formation of d-3 (Fig. 1). When the dehydrogenation was conducted in a J-Young tube in C6D6, a characteristic triplet corresponding to HD appeared at δ 4.43 (JHD = 45 Hz) in the 1H NMR spectrum (Fig. S18).36The structures of 2 and 3 were verified by X-ray crystallographic analysis (Fig. 2). Consistent with NMR spectroscopic analysis, the BH3 moiety in 2 is stabilized by one of the B–H bonds binding at the Fe–NH unit to form an Fe–H–B–N four-membered metallacycle. This metal–ligand cooperative binding mode increased the B–H bond length in the bridging B–H(1) bond to 1.362 Å vs. 1.129 Å and 1.121 Å for the two terminal B–H bonds. The B–N bond length of 1.545(3) Å in 2 is slightly shorter than that in H3B·NH3 (dB–N = 1.58(2) Å).37 Crystallographic analysis of 3 confirmed an imino-borane complex with a Cp*Fe(η2-H2B Created by potrace 1.16, written by Peter Selinger 2001-2019 NC6H4PPh2) framework. After dehydrogenation of 2, striking structural changes were observed. The N atom has been become detached from Fe, while the BH2 fragment acts as a bis(σ-borane) ligand coordinated to the metal center.21–23 The B–N bond distance of 1.455(5) Å in 3 is shorter by 0.09 Å than that in 2, and is close to that reported for the cyclic trimer borazine (1.4355(21) Å).38 Combined with the NMR results, the B–N bond length in 3 suggests some double bond character.21,22 As the imino-borane fragment is tethered in the coordination sphere, the boron center adopts a quasi-tetrahedral geometry, and the B–N bond appears to be partially sp3 hybridized. Dehydrogenation of the amido-borane complex also caused the decrease of the Fe⋯B distances from 2.223(3) Å to 2.026(4) Å which is shorter than the sum of the covalent radii of Fe and B atom (2.16 Å), indicating that the borane and the metal are bonded.Open in a separate windowFig. 2Solid-sate structure (50% probability thermal ellipsoids) of (a) complex 2 and (b) 3. For clarity, hydrogen atoms of Cp* and phenyl rings are omitted.Notably, the amido-borane compound 2 can be regenerated by stepwise protonation of 3 and transfer of a hydride (Scheme 3). Complex 3 reacts readily with H(Et2O)2BArF4 in C6H5F. The reaction solution was analyzed by ESI-MS spectroscopy, which showed an ionic peak at m/z = 480.1726 (calcd m/z = 480.1715), suggesting the formation of [3H]+. Alternatively, the reaction of complex 2 with H(Et2O)2BArF4 unambiguously provides [3H]+ and produces H2. X-ray crystallographic analysis reveals that the resulting cationic complex [3H]+ exhibits a similar framework to its imino-borane precursor (3). The BH2 moiety retains a binding mode of the bis(σ-BH2) fashion (Fig. 3). In contrast, the B–N distance in [3H]+ (1.586(6) Å) is extended by 0.13 Å and the [3H]+ framework becomes much less compact than that of 3. Probably due to the fluxional structure of the seven-membered Fe–P–C–C–N–B(H) ring, the solution of [3H][BArF4] gives broad 1H NMR resonances even at −60 °C. The phosphorus resonance arose at δ 72.0 as a singlet when the solution sample was cooled to −40 °C (Fig. S20 and S21).Open in a separate windowFig. 3Solid-state structures of (a) complex [3H]+ and (b) [3H(PPh3)]+. For clarity, counterion [BArF4], hydrogen atoms of Cp* and phenyl rings have been omitted.Open in a separate windowScheme 3Conversions of iron imino-borane, amino-borane and amido-borane complexes.In [3H]+, the boron is coordinatively unsaturated, as manifested by its interaction with a σ-donor. For instance, treatment of 2 with [HPPh3][BArF4] (pKMeCNa = 7.6)39 provides a Ph3P-stabilized borane complex, [3H(PPh3)]+ (m/z = 742.2620, calcd m/z = 742.2626). The 1H NMR spectrum of [3H(PPh3)]+ exhibits an NH resonance at δ 4.68, suggesting that protonation occurred at the N site. The distinctive upfield hydride signal for Fe–H–B is observed at δ −15.58. In the 31P NMR spectrum, two phosphorus signals at δ 78.90 and −1.26 correspond to the Fe–P and the B–P resonances, respectively. The 11B signal at δ −13.72 indicates a tetracoordinated boron, which is further confirmed by crystallographic analysis of [3H(PPh3)]+ (Fig. 3). In the solid-sate structure, a Ph3P molecule is bound to the B center (dB–P = 1.982(4) Å), leading to the formation of a new Fe–H–B–N four-membered metallacycle. As a amido-borane complex, [3H(PPh3)]+ has a B–N bond length of 1.527(5) Å, somewhat shorter than 1.545(3) Å in 2.After attaching a proton at the N atom, we subsequently explored restoration of the original borane moiety. Treatment of freshly prepared [3H][BArF4] in fluorobenzene with catecholborane-NEt3 adduct (δB = 10.56, JHB = 142.4 Hz)40 results in the regeneration of 2, as evidenced by the NMR spectra (Fig. S29 and S30). The 1H NMR spectrum of the reaction mixture displays a characteristic hydride signal at −13.97 ppm, indicating the recovery of the iron amido-borane complex. On the other side, concomitant formation of the borenium ion (δB = 13.86) was also observed in the 11B NMR spectrum, which agrees with the hydride transfer from the organohydride reagent to [3H]+. It was interesting that the ion [3H]+ is stable towards 5,6-dihydrophenanthridine and Hantszch ester. These results indicate that the hydride-donating ability (ΔGH) of 2 is in the range of 55–59 kcal mol−1.41 The reactive nature of the hydride in 2 was demonstrated by the reaction with [HPPh3][BArF4], which produces [3H(PPh3)]+ and releases H2 (Scheme 3).1The metal amine-borane complexes and their dehydrogenated derivatives are implicated throughout the catalytic cycle of amine-borane dehydrogenation. We found both the iron complexes 2 and 3 are efficient catalysts for H3N·BH3 dehydrogenation at room temperature. In the presence of 1 mol% catalyst, a THF solution of H3N·BH3 (1.0 mmol) generates about 2.2 equivalent of H2 within 6 h based on GC quantification (Fig. S33). More importantly, such catalytic dehydrocoupling systems allow for selective transfer hydrogenation of quinolines to dihydroquinolines, which are valuable synthons leading to many bio-active compounds.42 For instance, addition of methyl-6-quinolineacetate (4) to the catalytic system containing one equiv. of H3N·BH3 and 1 mol% of 3 gave 1,2-dihydro-methyl-6-quinolineacetate (5) in excellent yield within 6 h (eqn (1)). The outcome of this reaction was unaffected by switching the catalyst from 3 to 2, or by use of excess reducing agent or by an increase in the reaction temperature (Table S1).  相似文献   

14.
The degradation of neurotransmitters is a hallmark feature of Alzheimer''s disease (AD). Copper bound Aβ peptides, invoked to be involved in the pathology of AD, are found to catalyze the oxidation of serotonin (5-HT) by H2O2. A combination of EPR and resonance Raman spectroscopy reveals the formation of a Cu(ii)–OOH species and a dimeric, EPR silent, Cu2O2 bis-μ-oxo species under the reaction conditions. The Cu(ii)–OOH species, which can be selectively formed in the presence of excess H2O2, is the reactive intermediate responsible for 5-HT oxidation. H2O2 produced by the reaction of O2 with reduced Cu(i)–Aβ species can also oxidize 5-HT. Both these pathways are physiologically relevant and may be involved in the observed decay of neurotransmitters as observed in AD patients.

The mononuclear copper hydroperoxo species (Cu(ii)–OOH) of Cu–Aβ is the active oxidant responsible for serotonin oxidation by Cu–Aβ in the presence of physiologically relevant oxidants like O2 and H2O2, which can potentially cause oxidative degradation of neurotransmitters, a marker of Alzheimer''s disease.  相似文献   

15.
Sequential treatment of [Rh(COE)2Cl]2 (COE = cyclooctene) with PiPr3, alkyne derivatives and tBuN Created by potrace 1.16, written by Peter Selinger 2001-2019 BMes (Mes = 2,4,6-trimethylphenyl) provided functionalized rhodium η4-1,2-azaborete complexes of the form (η4-azaborete)RhCl(PiPr3). The scope of this reaction was expanded to encompass alkynes with hydrogen, alkyl, aryl, ferrocenyl, alkynyl, azaborinyl and boronate ester substituents. Treatment of these complexes with PMe3 led to insertion of the rhodium atom into the B–C bond of the BNC2 ring, forming 1-rhoda-3,2-azaboroles. Addition of N-heterocyclic carbenes to azaborete complexes led to highly unusual rearrangements to rhodium η21-allenylborylamino complexes via deprotonation and C–N bond cleavage. Heating and photolysis of an azaborete complex also led to rupture of the C–N bond followed by subsequent rearrangements, yielding an η4-aminoborylallene complex and two isomeric η4-butadiene complexes.

Rhodium η4-azaborete complexes can be transformed into a variety of species with ring-opened, BN-containing ligands by treatment with Lewis bases.  相似文献   

16.
Isocyanoazulenes (CNAz) constitute a relatively new class of isocyanoarenes that offers rich structural and electronic diversification of the organic isocyanide ligand platform. This article considers a series of 2-isocyano-1,3-X2-azulene ligands (X = H, Me, CO2Et, Br, and CN) and the corresponding zero-valent complexes thereof, [(OC)5Cr(2-isocyano-1,3-X2-azulene)]. Air- and thermally stable, X-ray structurally characterized 2-isocyano-1,3-dimethylazulene may be viewed as a non-benzenoid aromatic congener of 2,6-dimethyphenyl isocyanide (2,6-xylyl isocyanide), a longtime “workhorse” aryl isocyanide ligand in coordination chemistry. Single crystal X-ray crystallographic {Cr–CNAz bond distances}, cyclic voltametric {E1/2(Cr0/1+)}, 13C NMR {δ(13CN), δ(13CO)}, UV-vis {dπ(Cr) → pπ*(CNAz) Metal-to-Ligand Charge Transfer}, and FTIR {νNC, νCO, kCO} analyses of the [(OC)5Cr(2-isocyano-1,3-X2-azulene)] complexes provided a multifaceted, quantitative assessment of the π-acceptor/σ-donor characteristics of the above five 2-isocyanoazulenes. In particular, the following inverse linear relationships were documented: δ(13COtrans) vs. δ(13CN), δ(13COcis) vs. δ(13CN), and δ(13COtrans) vs. kCO,trans force constant. Remarkably, the net electron withdrawing capability of the 2-isocyano-1,3-dicyanoazulene ligand rivals those of perfluorinated isocyanides CNC6F5 and CNC2F3.  相似文献   

17.
The anisotropy barrier of polynuclear single-molecule magnets is expected to be higher with less tunneling the better stabilized the spin ground state is so that less MS mixing in the ground state and with excited spin states occur. We have realized this experimentally in two structurally related heptanuclear SMMs: the triplesalen-based [MnIII 6 CrIII]3+ and the triplesalalen-based *[MnIII 6 CrIII]3+ . The ligand system triplesalen was developed to enforce ferromagnetic interactions by the spin-polarization mechanism. However, we found weak antiferromagnetic couplings, that we assigned to an inefficient spin-polarization by a heteroradialene formation. To prevent this heteroradialene formation, the triplesalalen ligand H6talalen was designed. Here, we present the building block [(talalen )MnIII3]3+ and its application for the assembly of [{(talalen )MnIII3}2{CrIII(CN)6}]3+ (= *[MnIII 6 CrIII]3+ ). Both the trinuclear and heptanuclear complexes are SMMs. The comparison to the related triplesalen complex [(feld )MnIII3]3+ proves the absence of heteroradialene character and the enforcement of ferromagnetic MnIII-MnIII interactions in the (talalen )6− complexes. This results in an increase of the barrier for spin reversal Ueff from 25 K in the triplesalen-based [MnIII 6 CrIII]3+ SMMs to 37 K in the triplesalalen-based *[MnIII 6 CrIII]3+ SMM proving the success of our concept. Based on this study, the next step in the rational improvement of our SMMs is discussed.  相似文献   

18.
Subchalcogenides are uncommon, and their chemical bonding results from an interplay between metal–metal and metal–chalcogenide interactions. Herein, we present Ir6In32S21, a novel semiconducting subchalcogenide compound that crystallizes in a new structure type in the polar P31m space group, with unit cell parameters a = 13.9378(12) Å, c = 8.2316(8) Å, α = β = 90°, γ = 120°. The compound has a large band gap of 1.48(2) eV, and photoemission and Kelvin probe measurements corroborate this semiconducting behavior with a valence band maximum (VBM) of −4.95(5) eV, conduction band minimum of −3.47(5) eV, and a photoresponse shift of the Fermi level by ∼0.2 eV in the presence of white light. X-ray absorption spectroscopy shows absorption edges for In and Ir do not indicate clear oxidation states, suggesting that the numerous coordination environments of Ir6In32S21 make such assignments ambiguous. Electronic structure calculations confirm the semiconducting character with a nearly direct band gap, and electron localization function (ELF) analysis suggests that the origin of the gap is the result of electron transfer from the In atoms to the S 3p and Ir 5d orbitals. DFT calculations indicate that the average hole effective masses near the VBM (1.19me) are substantially smaller than the average electron masses near the CBM (2.51me), an unusual feature for most semiconductors. The crystal and electronic structure of Ir6In32S21, along with spectroscopic data, suggest that it is neither a true intermetallic nor a classical semiconductor, but somewhere in between those two extremes.

Subchalcogenides are uncommon, and their chemical bonding results from an interplay between metal–metal and metal–chalcogenide interactions.  相似文献   

19.
Despite their potential as promising alternatives to current state-of-the-art lithium-ion batteries, aqueous rechargeable Zn-ion batteries are still far away from practical applications. Here, we present a new class of single-ion conducting electrolytes based on a zinc sulfonated covalent organic framework (TpPa-SO3Zn0.5) to address this challenging issue. TpPa-SO3Zn0.5 is synthesised to exhibit single Zn2+ conduction behaviour via its delocalised sulfonates that are covalently tethered to directional pores and achieve structural robustness by its β-ketoenamine linkages. Driven by these structural and physicochemical features, TpPa-SO3Zn0.5 improves the redox reliability of the Zn metal anode and acts as an ionomeric buffer layer for stabilising the MnO2 cathode. Such improvements in the TpPa-SO3Zn0.5–electrode interfaces, along with the ion transport phenomena, enable aqueous Zn–MnO2 batteries to exhibit long-term cyclability, demonstrating the viability of COF-mediated electrolytes for Zn-ion batteries.

A zinc sulfonated covalent organic framework is presented as a new single-ion conducting electrolyte for aqueous rechargeable Zn-ion batteries.  相似文献   

20.
Uranium nitride compounds are important molecular analogues of uranium nitride materials such as UN and UN2 which are effective catalysts in the Haber–Bosch synthesis of ammonia, but the synthesis of molecular nitrides remains a challenge and studies of the reactivity and of the nature of the bonding are poorly developed. Here we report the synthesis of the first nitride bridged uranium complexes containing U(vi) and provide a unique comparison of reactivity and bonding in U(vi)/U(vi), U(vi)/U(v) and U(v)/U(v) systems. Oxidation of the U(v)/U(v) bis-nitride [K2{U(OSi(OtBu)3)3(μ-N)}2], 1, with mild oxidants yields the U(v)/U(vi) complexes [K{U(OSi(OtBu)3)3(μ-N)}2], 2 and [K2{U(OSi(OtBu)3)3}2(μ-N)2(μ-I)], 3 while oxidation with a stronger oxidant (“magic blue”) yields the U(vi)/U(vi) complex [{U(OSi(OtBu)3)3}2(μ-N)2(μ-thf)], 4. The three complexes show very different stability and reactivity, with N2 release observed for complex 4. Complex 2 undergoes hydrogenolysis to yield imido bridged [K2{U(OSi(OtBu)3)3(μ-NH)}2], 6 and rare amido bridged U(iv)/U(iv) complexes [{U(OSi(OtBu)3)3}2(μ-NH2)2(μ-thf)], 7 while no hydrogenolysis could be observed for 4. Both complexes 2 and 4 react with H+ to yield quantitatively NH4Cl, but only complex 2 reacts with CO and H2. Differences in reactivity can be related to significant differences in the U–N bonding. Computational studies show a delocalised bond across the U–N–U for 1 and 2, but an asymmetric bonding scheme is found for the U(vi)/U(vi) complex 4 which shows a U–N σ orbital well localised to U Created by potrace 1.16, written by Peter Selinger 2001-2019 N and π orbitals which partially delocalise to form the U–N single bond with the other uranium.

The first examples of molecular compounds containing the cyclic (U(vi)N)2 and (U(v)U(vi)N)2 cores were obtained by oxidation of the (U(v)U(v)N)2 analogue. Different bonding within these complexes yields different stability and reactivity with CO and H2.  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