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 421 毫秒
1.
By combining idealized experiments with realistic quantum mechanical simulations of an interface, we investigate electro-reduction reactions of HF, water and methanesulfonic acid (MSA) on the single crystal (111) facets of Au, Pt, Ir and Cu in organic aprotic electrolytes, 1 M LiPF6 in EC/EMC 3:7W (LP57), the aprotic electrolyte commonly used in Li-ion batteries, 1 M LiClO4 in EC/EMC 3:7W and 0.2 M TBAPF6 in 3 : 7 EC/EMC. In our previous work, we have established that LiF formation, accompanied by H2 evolution, is caused by a reduction of HF impurities and requires the presence of Li at the interface, which catalyzes the HF dissociation. In the present paper, we find that the measured potential of the electrochemical response for these reduction reactions correlates with the work function of the electrode surfaces and that the work function determines the potential for Li+ adsorption. The reaction path is investigated further by electrochemical simulations suggesting that the overpotential of the reaction is related to stabilizing the active structure of the interface having adsorbed Li+. Li+ is needed to facilitate the dissociation of HF which is the source of protons. Further experiments on other proton sources, water and methanesulfonic acid, show that if the hydrogen evolution involves negatively charged intermediates, F or HO, a cation at the interface can stabilize them and facilitate the reaction kinetics. When the proton source is already significantly dissociated (in the case of a strong acid), there is no negatively charged intermediate and thus the hydrogen evolution can proceed at much lower overpotentials. This reveals a situation where the overpotential for electrocatalysis is related to stabilizing the active structure of the interface, facilitating the reaction rather than providing the reaction energy.

By combining idealized experiments with realistic quantum mechanical simulations of an interface, we investigate electroreduction reactions of HF, water and methanesulfonic acid on the single crystal (111) facets of Au, Pt, Ir and Cu in a variety of aprotic electrolytes.  相似文献   

2.
Cyanuric triazide reacts with several transition metal precursors, extruding one equivalent of N2 and reducing the putative diazidotriazeneylnitrene species by two electrons, which rearranges to N-(1′H-[1,5′-bitetrazol]-5-yl)methanediiminate (biTzI2−) dianionic ligand, which ligates the metal and dimerizes, and is isolated from pyridine as [M(biTzI)]2Py6 (M = Mn, Fe, Zn, Cu, Ni). Reagent scope, product analysis, and quantum chemical calculations were combined to elucidate the mechanism of formation as a two-electron reduction preceding ligand rearrangement.

Cyanuric triazide reacts with transition metal precursors, extruding N2 and reducing the ligand by two electrons, which breaks an aromatic ring and rearranges to a bitetrazolylmethanediiminate (biTzI2−) ligand, forming two new aromatic rings.  相似文献   

3.
Here, we report the nitric oxide monooxygenation (NOM) reactions of a CoIII-nitrosyl complex (1, {Co-NO}8) in the presence of mono-oxygen reactive species, i.e., a base (OH, tetrabutylammonium hydroxide (TBAOH) or NaOH/15-crown-5), an oxide (O2− or Na2O/15-crown-5) and water (H2O). The reaction of 1 with OH produces a CoII-nitrito complex {3, (CoII-NO2)} and hydrogen gas (H2), via the formation of a putative N-bound Co-nitrous acid intermediate (2, {Co-NOOH}+). The homolytic cleavage of the O–H bond of proposed [Co-NOOH]+ releases H2via a presumed CoIII-H intermediate. In another reaction, 1 generates CoII-NO2 when reacted with O2−via an expected CoI-nitro (4) intermediate. However, complex 1 is found to be unreactive towards H2O. Mechanistic investigations using 15N-labeled-15NO and 2H-labeled-NaO2H (NaOD) evidently revealed that the N-atom in CoII-NO2 and the H-atom in H2 gas are derived from the nitrosyl ligand and OH moiety, respectively.

Base-induced hydrogen (H2) gas evolution in the nitric oxide monoxygenation reaction.

As a radical species, nitric oxide (NO) has attracted great interest from the scientific community due to its major role in various physiological processes such as neurotransmission, vascular regulation, platelet disaggregation and immune responses to multiple infections.1 Nitric oxide synthase (NOS),2 and nitrite reductase (NiR)3 enzymes are involved in the biosynthesis of NO. NOSs produce NO by the oxidation of the guanidine nitrogen in l-arginine.4 However, in mammals and bacteria, NO2 is reduced to NO by NiRs in the presence of protons, i.e., NO2 + e + 2H+ → NO + H2O.5 Biological dysfunctions may cause overproduction of NO, and being radical it leads to the generation of reactive nitrogen species (RNS), i.e., peroxynitrite (PN, OONO)6 and nitrogen dioxide (˙NO2),7 upon reaction with reactive oxygen species (ROS) such as superoxide (O2˙),8 peroxide (H2O2),9 and dioxygen (O2).10 Hence, it is essential to maintain an optimal level of NO. In this regard, nitric oxide dioxygenases (NODs)11 are available in bio-systems to convert excess NO to biologically benign nitrate (NO3).12NO2 + FeII + H+ ↔ NO + FeIII + OH1[M–NO]n + 2OH → [M–NO2](n−2) + H2O2NOD enzymes generate NO3 from NO;11b,12−13 however, the formation of NO2 from NO is still under investigation. Clarkson and Bosolo reported NO2 formation in the reaction of CoIII-NO and O2.14 Nam and co-workers showed the generation of CoII-NO2 from CoIII-NO upon reaction with O2˙.15 Recently, Mondal and co-workers reported NO2 formation in the reaction of CoII-NO with O2.16 Apart from cobalt, the formation of CuII-NO2 was also observed in the reaction of CuI-NO and O2.17 For metal-dioxygen adducts, i.e., CrIII-O2˙ and MnIV-O22−, NOD reactions led to the generation of CrIII-NO2 (ref. 18) and MnV Created by potrace 1.16, written by Peter Selinger 2001-2019 O + NO2,19 respectively. However, the NOD reaction of FeIII-O2˙ and FeIII-O22− with NO and NO+, respectively, generated FeIII-NO3via FeIV Created by potrace 1.16, written by Peter Selinger 2001-2019 O and ˙NO2.20 Ford suggested that the reaction of ferric-heme nitrosyl with hydroxide leads to the formation of NO2 and H+.12 Lehnert and co-workers reported heme-based Fe-nitrosyl complexes21 showing different chemistries due to the FeII-NO+ type electronic structures. On the other hand, Bryan proposed that the one-electron reduction of NO2 to NO in ferrous heme protein is reversible (eqn (1)).22 Also, it is proposed that excess NO in biological systems is converted to NO2 and produces one equivalent of H+ upon reaction with ˙OH.23 Previously reported reactivity of M–NOs of Fe24 with OH suggested the formation of NO2 and one equivalent of H+, where H+ further reacts with one equivalent of OH and produces H2O (eqn (2)).25Here in this report, we explore the mechanistic aspects of nitric oxide monooxygenation (NOM) reactions of the CoIII-nitrosyl complex, [(12TMC)CoIII(NO)]2+/{Co(NO)}8 (1),15,26 bearing the 12TMC ligand (12TMC = 1,4,7,10-tetramethyl-1,4,7,10-tetraazacyclododecane) with mono-oxygen reactive species (O2−, OH and H2O) (Scheme 1). Complex 1 reacts with the base (OH, tetrabutylammonium hydroxide (TBAOH)/or NaOH in the presence of 15-crown-5 as the OH source) and generates the corresponding CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), with the evolution of hydrogen gas (H2) via the formation of a plausible N-bound Co-nitrous acid intermediate ([Co-NOOH]+, 2) in CH3CN at 273 K (Scheme 1, reaction (I)). Also, when 1 reacts with the oxide (O2− or Na2O in the presence of 15-crown-5), it generates the CoII-nitrito complex (3) via a probable CoI-nitro, [(12TMC)CoI(NO2)] (4), intermediate (Scheme 1, reaction (II)); however, 1 does not react with water (Scheme 1, reaction (III)). Mechanistic investigations using 15N-labeled-15NO, D-labeled-NaOD and 18O-labelled-18OH demonstrated, unambiguously, that the N and O-atoms in the NO2 ligand of 3 resulted from NO and OH moieties; however, the H-atoms of H2 are derived from OH. To the extent of our knowledge, the present work reports the very first systematic study of CoIII-nitrosyl complex reactions with H2O, OH and O2−. This new finding presents an alternative route for NO2 generation in biosystems, and also illustrates a new pathway of H2 evolution, in addition to the reported literature.12,27Open in a separate windowScheme 1Nitric oxide monooxygenation (NOM) reactions of cobalt-nitrosyl complex (1) in the presence of a base (OH), sodium oxide (Na2O) and water (H2O).To further explore the chemistry of [(12TMC)CoIII(NO)]2+ (1),15,26 and the mechanistic insights of NOM reactions, we have reacted it with a base (OH), an oxide (O2−), and water (H2O). When complex 1 was reacted with TBAOH in CH3CN, the color of complex 1 changed to light pink from dark pink. In this reaction, the characteristic absorption band of 1 (370 nm) disappears within 2 minutes (Fig. 1a; ESI, Experimental section (ES) and Fig. S1a), producing a CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), with H2 (Scheme 1, reaction (Ib)), in contrast to the previous reports on base induced NOM reactions (eqn (2)).12,25,28 The spectral titration data confirmed that the ratio-metric equivalent of OH to 1 was 1 : 1 (ESI, Fig. S1b). 3 was determined to be [(12TMC)CoII(NO2)](BF4) based on various spectroscopic and structural characterization experiments (vide infra).15,26bOpen in a separate windowFig. 1(a) UV-vis spectral changes of 1 (0.50 mM, black line) upon addition of OH (1 equiv.) in CH3CN under Ar at 273 K. Black line (1) changed to red line (3) upon addition of OH. Inset: IR spectra of 3-14NO2 (blue line) and 3-15NO2 (red line) in KBr. (b) ESI-MS spectra of 3. The peak at 333.2 is assigned to [(12TMC)CoII(NO2)]+ (calcd m/z 333.1). Inset: isotopic distribution pattern for 3-14NO2 (red line) and 3-15NO2 (blue line).The FT-IR spectrum of 3 showed a characteristic peak for nitrite stretching at 1271 cm−1 (CoII-14NO2) and shifted to 1245 cm−1 (CoII-15NO2) when 3 was prepared by reacting 15N-labeled NO (CoIII-15NO) with OH (Inset, Fig. 1a and Fig. S2). The shifting of NO2 stretching (Δ = 30 cm−1) indicates that the N-atom in the NO2 ligand is derived from CoIII-15NO. The ESI-MS spectrum of 3 showed a prominent peak at m/z 333.2, [(12TMC)CoII(14NO2)]+ (calcd m/z 333.2), which shifted to 334.2, [(12TMC)CoII(15NO2)]+ (calcd m/z 334.2), when the reaction was performed with CoIII-15NO (Inset, Fig. 1b; ESI, Fig. S3a); indicating clearly that NO2 in 3 was derived from the NO moiety of 1. In addition, we have reacted 1 with Na18OH (ES and ESI), in order to follow the source of the second O-atom in 3-NO2. The ESI-MS spectrum of the reaction mixture, obtained by reacting 1 with Na18OH, showed a prominent peak at m/z 335.2, [(12TMC)CoII(18ONO)]+ (calcd m/z 335.2), (SI, Fig. S3b) indicating clearly that NO2 in 3 was derived from 18OH. The 1H NMR spectrum of 3 did not show any signal for aliphatic protons of the 12TMC ligand, suggesting a bivalent cobalt center (Fig. S4).26b Furthermore, we have determined the magnetic moment of 3, using Evans'' method, and it was found to be 4.62 BM, suggesting a high spin Co(ii) metal center with three unpaired electrons (ESI and ES).29 The exact conformation of 3 was provided by single-crystal X-ray crystallographic analysis (Fig. 2b, ESI, ES, Fig. S5, and Tables T1 and T2) and similar to that of previously reported CoII-NO2/MII-NO2.15,26b Also, we have quantified the amount of nitrite (90 ± 5%), formed in the above reaction, using the Griess reagent (ESI, ES, and Fig. S6).Open in a separate windowFig. 2Displacement ellipsoid plot (20% probability) of 3 at 100 K. Disordered C-atoms of the TMC ring, anion and H-atoms have been removed for clarity.As is known from the literature, a metal-nitrous acid intermediate may form either by the reaction of a metal-nitrosyl with a base27 or by the metal-nitrite reaction with an acid (nitrite reduction chemistry);26b however, the products of both the reactions are different. Here, for the first time, we have explored the reaction of CoIII-nitrosyl (1) with a base. In this reaction, it is clear that the formation of CoII-nitrito would be accomplished by the release of H2 gas via the generation of a transient N-bound [Co-(NOOH)]+ intermediate (Scheme 2, reaction (II)). The formation of CoII-NO2 (3) from the [Co-(NOOH)]+ intermediate is likely to proceed by either (i) homolytic cleavage of the O–H bond and release of H2via the proposed CoIII-H transient species (CoIII-H = CoII + 1/2H2)30 (Scheme 2, reaction (III)), as reported in previous literature where the reduced cobalt, in a number of different ligand environments, is a good H+ reduction catalyst and generates H2 gas via a CoIII-H intermediate31 or (ii) heterolytic cleavage of the O–H bond and the formation of CoI-NO2 + H+.27 In the present study, we observed the formation of 3 and H2via the plausible homolytic cleavage of the NOO–H moiety of 2 as shown in Scheme 2, in contrast to the previous reports on base-induced reactions on metal-nitrosyls (eqn (3)).27 Taking together both possibilities, (i) is the most reasonable pathway for the NOM reaction of complex 1 in the presence of a base (as shown in Scheme 2, reaction (III)). And the reaction is believed to go through a CoIII-H intermediate as reported previously in CoI-induced H+ reduction in different ligand frameworks and based on literature precedence, we believe that complex 1 acts in a similar manner.31Open in a separate windowScheme 2NOM reaction of complex 1 in the presence of OH, showing the generation of CoII-nitrito (3) and H2via a Co(iii)-hydrido intermediate.In contrast to an O-bound CoII-ONOH intermediate, where N–O bond homolysis of the ON-OH moiety generates H2O2 (Scheme 2, reaction (IV)),26b the N-bound [Co-(NOOH)]+ intermediate decomposes to form NO2 and a Co(iii)-H transient species, arising from β-hydrogen transfer from the NOO–H moiety to the cobalt-center (Scheme 2, reaction (II)).30a,c,32 The Co(iii)-hydrido species may generate H2 gas either (a) by its transformation to the Co(ii)-nitrito complex (2) and H2 gas as observed in the case of CoIII-H intermediate chemistry30a,c,e−g as proposed in the chemistry of the CoI complex with H+ reduction31 and other metal-hydrido intermediates32 and also explained in O2 formation in PN chemistry17,33 or (b) by the reacting with another [Co-(NOOH)]+ intermediate (Scheme 2, reaction (III)).Furthermore, we have confirmed the H2 formation in the NOM reaction of 1 with OH by headspace gas mass spectrometry (Fig. 3a). Also, carrying out the reaction of 1 with NaOD leads to the formation of the [Co-(NOOD)]+ intermediate, which then transforms to a CoIII-D transient species. Further, as described above, the CoIII-D species releases D2 gas, detected by headspace gas mass spectrometry (Fig. 3b), which evidently established that H2 gas formed in the reaction of 1 with OH. In this regard, we have proposed that in the first step of this reaction, the nucleophilic addition of OH to {Co-NO}8 generates a transient N-bound [Co-(NOOH)]+ intermediate that is generated by an internal electron transfer to CoIII (Scheme 2, reaction (I)). By following the mechanism proposed in the case of CoIII-H,30a−c O2,15 and H2O2(ref. 26b) formation, we have proposed the sequences of the NOM reaction of 1, which leads to the generation of CoII-nitrito and H2 (Scheme 2, reaction (I)–(III) and Scheme 3). In the second step, O–H bond homolytic cleavage generates a CoIII-H transient species + NO2via a β-hydrogen elimination reaction of the [Co-(NOOH)]+ intermediate.32 The CoIII-H intermediate may undergo the following reactions to generate H2 gas and CoII-nitrito either (a) by the natural decomposition of the CoIII-H transient species to generate H2,30a,c,e−g or (b) by the H-atom abstraction from another [Co-(NOOH)]+ intermediate (Scheme 3). Also, to validate our assumption that the reaction goes through a plausible N-bound [Co-(NOOH)]+ intermediate followed by its transformation to the CoIII-H species (vide supra), we have performed the reaction of 1 with NaOH/NaOD (in 1 : 1 ratio). In this reaction, we have observed the formation of a mixture of H2, D2, and HD gases, which indicates clearly that the reaction goes through the formation of CoIII-H and CoIII-D transient species via the aforementioned mechanism (Fig. 3c). This is the only example where tracking of the H atoms has confirmed the H2 generation from an N-bound NOO–H moiety as proposed for H2 formation from CoIII-H.30Open in a separate windowFig. 3Mass spectra of formation of (a) H2 in the reaction of 1 (5.0 mM) with NaOH (5.0 mM), (b) D2 in the reaction of 1 (5.0 mM) with NaOD (5.0 mM), (c) D2, HD, and H2 in the reaction of 1 (5.0 mM) with NaOD/NaOH (1 : 1), and (d) H2 in the reaction of 1 (5.0 mM) with NaOH in the presence of 2,4 DTBP (50 mM).Open in a separate windowScheme 3NOM reaction of complex 1 in the presence of OH, showing the different steps of the reaction.While, we do not have direct spectral evidence to support the formation of the transient N-bound [Co-(NOOH)]+ intermediate and its decomposition to the CoIII-H transient species via β-hydrogen transfer from the NOOH moiety to the cobalt center, support for its formation comes from our finding that the reactive hydrogen species can be trapped by using 2,4-di-tert-butyl-phenol (2,4-DTBP).34 In this reaction, we observed the formation of 2,4-DTBP-dimer (2,4-DTBP-D, ∼67%) as a single product (ESI, ES, and Fig. S7). This result can readily be explained by the H-atom abstraction reaction of 2,4-DTBP either by [Co-(NOOH)]+ or CoIII-H, hence generating a phenoxyl-radical and 3 with H2 (Fig. 3d and Scheme 2, reaction (a)). Also, we have detected H2 gas formation in this reaction (ESI, ES, and Fig. 3d). In the next step, two phenoxyl radicals dimerized to give 2,4-DTBP-dimer (Scheme 2c, reaction (II)). Thus, the observation of 2,4-DTBP-dimer in good yield supports the proposed reaction mechanism (Scheme 2, reaction (a) and (b)). Further, the formation of 2,4 DTBP as a single product also rules out the formation of the hydroxyl radical as observed in the case of an O-bound nitrous acid intermediate.26bFurthermore, we have explored the NOM reactivity of 1 with Na2O/15-crown-5 (as the O2− source) and observed the formation of the CoII-nitrito complex (3) via a plausible CoI-nitro (4) intermediate (Scheme 1, reaction (IIa); also see the ESI and ES); however, 1 was found to be inert towards H2O (Scheme 1, reaction (III); also see the ESI, ES and Fig. S8). The product obtained in the reaction of 1 with O2− was characterized by various spectroscopic measurements.15,26b The UV-vis absorption band of 1 (λmax = 370 nm) disappears upon the addition of 1 equiv. of Na2O and a new band (λmax = 535 nm) forms, which corresponds to 3 (ESI, Fig. S9). The FT-IR spectrum of the isolated product of the above reaction shows a characteristic peak for CoII-bound nitrite at 1271 cm−1, which shifts to 1245 cm−1 when exchanged with 15N-labeled-NO (15N16O) (ESI, ES, and Fig. S10), clearly indicating the generation of nitrite from the NO ligand of complex 1.26b The ESI-MS spectrum recorded for the isolated product (vide supra) shows a prominent ion peak at m/z 333.1, and its mass and isotope distribution pattern matches with [(12-TMC)CoII(NO2)]+ (calc. m/z 333.1) (ESI, Fig. S11). Also, we quantified the amount of 3 (85 ± 5%) by quantifying the amount of nitrite (85 ± 5%) using the Griess reagent test (ESI, ES, and Fig. S6).In summary, we have demonstrated the reaction of CoIII-nitrosyl, [(12-TMC)CoIII(NO)]2+/{CoNO}8 (1), with mono-oxygen reactive species (O2−, OH and H2O) (Scheme 1). For the first time, we have established the clear formation of a CoII-nitrito complex, [(12TMC)CoII(NO2)]+ (3), and H2 in the reaction of 1 with one equivalent of OHvia a transient N-bound [Co-(NOOH)]+ (2) intermediate. This [Co-(NOOH)]+ intermediate undergoes the O–H bond homolytic cleavage and generates a CoIII-H transient species with NO2, via a β-hydrogen elimination reaction of the [Co-(NOOH)]+ intermediate, which upon decomposition produces H2 gas. This is in contrast to our previous report, where acid-induced nitrite reduction of 3 generated 1 and H2O2via an O-bound CoII-ONOH intermediate.26b Complex 1 was found to be inert towards H2O; however, we have observed the formation of 3 when reacted with O2−. It is important to note that H2 formation involves a distinctive pathway of O–H bond homolytic cleavage in the [Co-(NOOH)]+ intermediate, followed by the generation of the proposed CoIII-H transient species (CoII + 1/2H2)30 prior to H2 evolution as described in CoI chemistry with H+ in many different ligand frameworks.31 The present study is the first-ever report where the base induced NOM reaction of CoIII-nitrosyl (1) leads to CoII-nitrito (3) with H2 evolution via an N-bound [Co-(NOOH)]+ intermediate, in contrast to the chemistry of O-bound CoII-ONOH26b, hence adding an entirely new mechanistic insight of base induced H2 gas evolution and an additional pathway for NOM reactions.  相似文献   

4.
5.
It was shown that chromium(VI) can be determined by coulometry at a Pt electrode in HNO3and H2SO4solutions. A procedure for the determination of 1–4 mg of chromium in 0.5–1 M HNO3solutions by controlled-potential coulometry with RSD = 0.2% was developed. It was demonstrated that the degree of the reduction of Cr(VI) to Cr(III) reached 99% (RSD = 0.5%) in the determination of chromium by coulometry at a slow potential sweep (v= 1–2 mV/s) in HNO3solutions. A procedure was proposed for the determination of Cr(VI) and Cu(II) simultaneously present in 0.25–0.5 M H2SO4solutions by controlled-potential coulometry with RSD = 0.3%.  相似文献   

6.
Designing solid-state electrolytes for proton batteries at moderate temperatures is challenging as most solid-state proton conductors suffer from poor moldability and thermal stability. Crystal–glass transformation of coordination polymers (CPs) and metal–organic frameworks (MOFs) via melt-quenching offers diverse accessibility to unique properties as well as processing abilities. Here, we synthesized a glassy-state CP, [Zn3(H2PO4)6(H2O)3](1,2,3-benzotriazole), that exhibited a low melting temperature (114 °C) and a high anhydrous single-ion proton conductivity (8.0 × 10−3 S cm−1 at 120 °C). Converting crystalline CPs to their glassy-state counterparts via melt-quenching not only initiated an isotropic disordered domain that enhanced H+ dynamics, but also generated an immersive interface that was beneficial for solid electrolyte applications. Finally, we demonstrated the first example of a rechargeable all-solid-state H+ battery utilizing the new glassy-state CP, which exhibited a wide operating-temperature range of 25 to 110 °C.

Melt-quenched coordination polymer glass shows exclusive H+ conductivity (8.0 × 10−3 S cm−1 at 120 °C, anhydrous) and optimal mechanical properties (42.8 Pa s at 120 °C), enables the operation of an all-solid-state proton battery from RT to 110 °C.  相似文献   

7.
Electrocatalytic synthesis of multicarbon (C2+) products from CO2 reduction suffers from poor selectivity and low energy efficiency. Herein, a facile oxidation–reduction cycling method is adopted to reconstruct the Cu electrode surface with the help of halide anions. The surface composed of entangled Cu nanowires with hierarchical pores is synthesized in the presence of I, exhibiting a C2 faradaic efficiency (FE) of 80% at −1.09 V vs. RHE. A partial current density of 21 mA cm−2 is achieved with a C2 half-cell power conversion efficiency (PCE) of 39% on this electrode. Such high selective C2 production is found to mainly originate from CO intermediate enrichment inside hierarchical pores rather than the surface lattice effect of the Cu electrode.

The Cu electrode surface is reconstructed by a halide anion assisted method for promoting CO2 reduction.  相似文献   

8.
9.
The use of radical bridging ligands to facilitate strong magnetic exchange between paramagnetic metal centers represents a key step toward the realization of single-molecule magnets with high operating temperatures. Moreover, bridging ligands that allow the incorporation of high-anisotropy metal ions are particularly advantageous. Toward these ends, we report the synthesis and detailed characterization of the dinuclear hydroquinone-bridged complexes [(Me6tren)2MII2(C6H4O22−)]2+ (Me6tren = tris(2-dimethylaminoethyl)amine; M = Fe, Co, Ni) and their one-electron-oxidized, semiquinone-bridged analogues [(Me6tren)2MII2(C6H4O2˙)]3+. Single-crystal X-ray diffraction shows that the Me6tren ligand restrains the metal centers in a trigonal bipyramidal geometry, and coordination of the bridging hydro- or semiquinone ligand results in a parallel alignment of the three-fold axes. We quantify the p-benzosemiquinone–transition metal magnetic exchange coupling for the first time and find that the nickel(ii) complex exhibits a substantial J < −600 cm−1, resulting in a well-isolated S = 3/2 ground state even as high as 300 K. The iron and cobalt complexes feature metal–semiquinone exchange constants of J = −144(1) and −252(2) cm−1, respectively, which are substantially larger in magnitude than those reported for related bis(bidentate) semiquinoid complexes. Finally, the semiquinone-bridged cobalt and nickel complexes exhibit field-induced slow magnetic relaxation, with relaxation barriers of Ueff = 22 and 46 cm−1, respectively. Remarkably, the Orbach relaxation observed for the Ni complex is in stark contrast to the fast processes that dominate relaxation in related mononuclear NiII complexes, thus demonstrating that strong magnetic coupling can engender slow magnetic relaxation.

A semiquinone radical bridging two trigonal bipyramidal metal centers facilitates strong magnetic exchange and single-molecule magnet behavior.  相似文献   

10.
An efficient strategy for designing charge-transfer complexes using coinage metal cyclic trinuclear complexes (CTCs) is described herein. Due to opposite quadrupolar electrostatic contributions from metal ions and ligand substituents, [Au(μ-Pz-(i-C3H7)2)]3·[Ag(μ-Tz-(n-C3F7)2)]3 (Pz = pyrazolate, Tz = triazolate) has been obtained and its structure verified by single crystal X-ray diffraction – representing the 1st crystallographically-verified stacked adduct of monovalent coinage metal CTCs. Abundant supramolecular interactions with aggregate covalent bonding strength arise from a combination of M–M′ (Au → Ag), metal–π, π–π interactions and hydrogen bonding in this charge-transfer complex, according to density functional theory analyses, yielding a computed binding energy of 66 kcal mol−1 between the two trimer moieties – a large value for intermolecular interactions between adjacent d10 centres (nearly doubling the value for a recently-claimed Au(i) → Cu(i) polar-covalent bond: Proc. Natl. Acad. Sci. U.S.A., 2017, 114, E5042) – which becomes 87 kcal mol−1 with benzene stacking. Surprisingly, DFT analysis suggests that: (a) some other literature precedents should have attained a stacked product akin to the one herein, with similar or even higher binding energy; and (b) a high overall intertrimer bonding energy by inferior electrostatic assistance, underscoring genuine orbital overlap between M and M′ frontier molecular orbitals in such polar-covalent M–M′ bonds in this family of molecules. The Au → Ag bonding is reminiscent of classical Werner-type coordinate-covalent bonds such as H3N: → Ag in [Ag(NH3)2]+, as demonstrated herein quantitatively. Solid-state and molecular modeling illustrate electron flow from the π-basic gold trimer to the π-acidic silver trimer with augmented contributions from ligand-to-ligand’ (LL′CT) and metal-to-ligand (MLCT) charge transfer.

A stacked Ag3–Au3 bonded (66 kcal mol−1) complex obtained crystallographically exhibits charge-transfer characteristics arising from multiple cooperative supramolecular interactions.  相似文献   

11.
In order to efficiently remove phosphorus, thermodynamic equilibrium diagrams of the P-H2O system and P-M-H2O system (M stands for Fe, Al, Ca, Mg) were analyzed by software from Visual MINTEQ to identify the existence of phosphorus ions and metal ions as pH ranged from 1 to 14. The results showed that the phosphorus ions existed in the form of H3PO4, H2PO4, HPO42−, and PO43−. Among them, H2PO4 and HPO42− were the main species in the acidic medium (99% at pH = 5) and alkaline medium (97.9% at pH = 10). In the P-Fe-H2O system ((P) = 0.01 mol/L, (Fe3+) = 0.01 mol/L), H2PO4 was transformed to FeHPO4+ at pH = 0–7 due to the existence of Fe3+ and then transformed to HPO42− at pH > 6 as the Fe3+ was mostly precipitated. In the P-Ca-H2O system ((P) = 0.01 mol/L, (Ca2+) = 0.015 mol/L), the main species in the acidic medium was CaH2PO4+ and HPO42−, and then transformed to CaPO4 at pH > 7. In the P-Mg-H2O system ((P) = 0.01 mol/L, (Mg2+) = 0.015 mol/L), the main species in the acidic medium was H2PO4 and then transformed to MgHPO4 at pH = 5–10, and finally transformed to MgPO4 as pH increased. The verification experiments (precipitation experiments) with single metal ions confirmed that the theoretical analysis could be used to guide the actual experiments.  相似文献   

12.
We report two macrocyclic ligands based on a 1,7-diaza-12-crown-4 platform functionalized with acetate (tO2DO2A2−) or piperidineacetamide (tO2DO2AMPip) pendant arms and a detailed characterization of the corresponding Mn(II) complexes. The X−ray structure of [Mn(tO2DO2A)(H2O)]·2H2O shows that the metal ion is coordinated by six donor atoms of the macrocyclic ligand and one water molecule, to result in seven-coordination. The Cu(II) analogue presents a distorted octahedral coordination environment. The protonation constants of the ligands and the stability constants of the complexes formed with Mn(II) and other biologically relevant metal ions (Mg(II), Ca(II), Cu(II) and Zn(II)) were determined using potentiometric titrations (I = 0.15 M NaCl, T = 25 °C). The conditional stabilities of Mn(II) complexes at pH 7.4 are comparable to those reported for the cyclen-based tDO2A2− ligand. The dissociation of the Mn(II) chelates were investigated by evaluating the rate constants of metal exchange reactions with Cu(II) under acidic conditions (I = 0.15 M NaCl, T = 25 °C). Dissociation of the [Mn(tO2DO2A)(H2O)] complex occurs through both proton− and metal−assisted pathways, while the [Mn(tO2DO2AMPip)(H2O)] analogue dissociates through spontaneous and proton-assisted mechanisms. The Mn(II) complex of tO2DO2A2− is remarkably inert with respect to its dissociation, while the amide analogue is significantly more labile. The presence of a water molecule coordinated to Mn(II) imparts relatively high relaxivities to the complexes. The parameters determining this key property were investigated using 17O NMR (Nuclear Magnetic Resonance) transverse relaxation rates and 1H nuclear magnetic relaxation dispersion (NMRD) profiles.  相似文献   

13.
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   

14.
Ligand-based mixed valent (MV) complexes of Al(iii) incorporating electron donating (ED) and electron withdrawing (EW) substituents on bis(imino)pyridine ligands (I2P) have been prepared. The MV states containing EW groups are both assigned as Class II/III, and those with ED functional groups are Class III and Class II/III in the (I2P)(I2P2−)Al and [(I2P2−)(I2P3−)Al]2− charge states, respectively. No abrupt changes in delocalization are observed with ED and EW groups and from this we infer that ligand and metal valence p-orbitals are well-matched in energy and the absence of LMCT and MLCT bands supports the delocalized electronic structures. The MV ligand charge states (I2P)(I2P2−)Al and [(I2P2−)(I2P3−)Al]2− show intervalence charge transfer (IVCT) transitions in the regions 6850–7740 and 7410–9780 cm−1, respectively. Alkali metal cations in solution had no effect on the IVCT bands of [(I2P2−)(I2P3−)Al]2− complexes containing –PhNMe2 or –PhF5 substituents. Minor localization of charge in [(I2P2−)(I2P3−)Al]2− was observed when –PhOMe substituents are included.

Organo-aluminum mixed-valent complexes combine properties of both organic and transition element mixed-valent compounds. This supports delocalized electronic structures that are structurally and electronically tunable.  相似文献   

15.
The reaction of [Co(CO)4] (1) with M(I) compounds (M = Cu, Ag, Au) was reinvestigated unraveling an unprecedented case of polymerization isomerism. Thus, as previously reported, the trinuclear clusters [M{Co(CO)4}2] (M = Cu, 2; Ag, 3; Au, 4) were obtained by reacting 1 with M(I) in a 2:1 molar ratio. Their molecular structures were corroborated by single-crystal X-ray diffraction (SC-XRD) on isomorphous [NEt4][M{Co(CO)4}2] salts. [NEt4](3)represented the first structural characterization of 3. More interestingly, changing the crystallization conditions of solutions of 3, the hexanuclear cluster [Ag2{Co(CO)4}4]2− (5) was obtained in the solid state instead of 3. Its molecular structure was determined by SC-XRD as Na2(5)·C4H6O2, [PPN]2(5)·C5H12 (PPN = N(PPh3)2]+), [NBu4]2(5) and [NMe4]2(5) salts. 5 may be viewed as a dimer of 3 and, thus, it represents a rare case of polymerization isomerism (that is, two compounds having the same elemental composition but different molecular weights) in cluster chemistry. The phenomenon was further studied in solution by IR and ESI-MS measurements and theoretically investigated by computational methods. Both experimental evidence and density functional theory (DFT) calculations clearly pointed out that the dimerization process occurs in the solid state only in the case of Ag, whereas Cu and Au related species exist only as monomers.  相似文献   

16.
H2S is a well-known toxic gas and also a gaseous signaling molecule involved in many biological processes. Advanced chemical tools that can regulate H2S levels in vivo are useful for understanding H2S biology as well as its potential therapeutic effects. To this end, we have developed a series of 7-nitro-1,2,3-benzoxadiazole (NBD) amines as potential H2S scavengers. The kinetic studies of thiolysis reactions revealed that incorporation of positively-charged groups onto the NBD amines greatly increased the rate of the H2S-specific thiolysis reaction. We demonstrate that these reactions proceed effectively, with second order rate constants (k2) of >116 M−1 s−1 at 37 °C for NBD-S8. Additionally, we demonstrate that NBD-S8 can effectively scavenge enzymatically-produced and endogenous H2S in live cells. Furthering the biological significance, we demonstrate NBD-S8 mediates scavenging of H2S in mice.

We demonstrate that positively-charged NBD amines can effectively scavenge biological H2S in live cells and in mice.  相似文献   

17.
Loading Ag and Co dual cocatalysts on Al-doped SrTiO3 (AgCo/Al-SrTiO3) led to a significantly improved CO-formation rate and extremely high selectivity toward CO evolution (99.8%) using H2O as an electron donor when irradiated with light at wavelengths above 300 nm. Furthermore, the CO-formation rate over AgCo/Al-SrTiO3 (52.7 μmol h−1) was a dozen times higher than that over Ag/Al-SrTiO3 (4.7 μmol h−1). The apparent quantum efficiency for CO evolution over AgCo/Al-SrTiO3 was about 0.03% when photoirradiated at a wavelength at 365 nm, with a CO-evolution selectivity of 98.6% (7.4 μmol h−1). The Ag and Co cocatalysts were found to function as reduction and oxidation sites for promoting the generation of CO and O2, respectively, on the Al-SrTiO3 surface.

Deposition Ag and Co dual cocatalysts onto Al-SrTiO3 significantly improves its activity for photoreduction of CO2 by H2O, with extremely high selectivity to CO evolution (99.8%), in which Ag and Co enable CO2 reduction and H2O oxidation, respectively.  相似文献   

18.
The self-assembly of 2,6-diformyl-4-methylphenol (DFMP) and 1-amino-2-propanol (AP)/2-amino-1,3-propanediol (APD) in the presence of copper(II) ions results in the formation of six new supramolecular architectures containing two versatile double Schiff base ligands (H3L and H5L1) with one-, two-, or three-dimensional structures involving diverse nuclearities: tetranuclear [Cu4(HL2−)2(N3)4]·4CH3OH·56H2O (1) and [Cu4(L3−)2(OH)2(H2O)2] (2), dinuclear [Cu2(H3L12−)(N3)(H2O)(NO3)] (3), polynuclear {[Cu2(H3L12−)(H2O)(BF4)(N3)]·H2O}n (4), heptanuclear [Cu7(H3L12−)2(O)2(C6H5CO2)6]·6CH3OH·44H2O (5), and decanuclear [Cu10(H3L12−)4(O)2(OH)2(C6H5CO2)4] (C6H5CO2)2·20H2O (6). X-ray studies have revealed that the basic building block in 1, 3, and 4 is comprised of two copper centers bridged through one μ-phenolate oxygen atom from HL2− or H3L12−, and one μ-1,1-azido (N3) ion and in 2, 5, and 6 by μ-phenoxide oxygen of L3− or H3L12− and μ-O2− or μ3-O2− ions. H-bonding involving coordinated/uncoordinated hydroxy groups of the ligands generates fascinating supramolecular architectures with 1D-single chains (1 and 6), 2D-sheets (3), and 3D-structures (4). In 5, benzoate ions display four different coordination modes, which, in our opinion, is unprecedented and constitutes a new discovery. In 1, 3, and 5, Cu(II) ions in [Cu2] units are antiferromagnetically coupled, with J ranging from −177 to −278 cm−1.  相似文献   

19.
Designing of porous carbon system for CO2 uptake has attracted a plenty of interest due to the ever-increasing concerns about climate change and global warming. Herein, a novel N rich porous carbon is prepared by in-situ chemical oxidation polyaniline (PANI) on a surface of multi-walled carbon nanotubes (MWCNTs), and then activated with KOH. The porosity of such carbon materials can be tuned by rational introduction of MWCNTs, adjusting the amount of KOH, and controlling the pyrolysis temperature. The obtained M/P-0.1-600-2 adsorbent possesses a high surface area of 1017 m2 g−1 and a high N content of 3.11 at%. Such M/P-0.1-600-2 adsorbent delivers an enhanced CO2 capture capability of 2.63 mmol g−1 at 298.15 K and five bars, which is 14 times higher than that of pristine MWCNTs (0.18 mmol g−1). In addition, such M/P-0.1-600-2 adsorbent performs with a good stability, with almost no decay in a successive five adsorption-desorption cycles.  相似文献   

20.
We report the synthesis of high-entropy-alloy (HEA) nanoparticles (NPs) consisting of five platinum group metals (Ru, Rh, Pd, Ir and Pt) through a facile one-pot polyol process. We investigated the electronic structure of HEA NPs using hard X-ray photoelectron spectroscopy, which is the first direct observation of the electronic structure of HEA NPs. Significantly, the HEA NPs possessed a broad valence band spectrum without any obvious peaks. This implies that the HEA NPs have random atomic configurations leading to a variety of local electronic structures. We examined the hydrogen evolution reaction (HER) and observed a remarkably high HER activity on HEA NPs. At an overpotential of 25 mV, the turnover frequencies of HEA NPs were 9.5 and 7.8 times higher than those of a commercial Pt catalyst in 0.05 M H2SO4 and 1.0 M KOH electrolytes, respectively. Moreover, the HEA NPs showed almost no loss during a cycling test and were much more stable than the commercial Pt catalyst. Our findings on HEA NPs may provide a new paradigm for the design of catalysts.

RuRhPdIrPt high-entropy-alloy nanoparticles with a broad and featureless valence band spectrum show high hydrogen evolution reaction activity.  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