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 718 毫秒
1.
A series of octanuclear iodine-bromine interhalides [InBr8−n]2− (n=0, 2, 3, 4) were prepared systematically in two steps. Firstly, addition of a dihalogen (Br2 or IBr) to the triaminocyclopropenium bromide salt [C3(NEt2)3]Br forms the corresponding trihalide salt with Br3 or IBr2 anions, respectively. Secondly, addition to Br3 of half an equivalent of Br2 gives the octabromine polyhalide [Br8]2−, whereas addition to IBr2 of half an equivalent of Br2, IBr or I2 gives the corresponding interhalides: [I2Br6]2−, [I3Br5]2−, and [I4Br4]2−, respectively. The four octahalides were characterized by X-ray crystallography, computational studies, Raman and Far-IR spectroscopies, as well as by TGA and melting point. All of the salts were found to be ionic liquids.  相似文献   

2.
The kinetics of the title reactions have been studied using the discharge-flow mass spectrometic method at 296 K and 1 torr of helium. The rate constant obtained for the forward reaction Br+IBr→I+Br2 (1), using three different experimental approaches (kinetics of Br consumption in excess of IBr, IBr consumption in excess of Br, and I formation), is: k1=(2.7±0.4)×10−11 cm3 molecule−1s−1. The rate constant of the reverse reaction: I+Br2→Br+IBr (−1) has been obtained from the Br2 consumption rate (with an excess of I atoms) and the IBr formation rate: k−1=(1.65±0.2)×10−13 cm3molecule−1s−1. The equilibrium constant for the reactions (1,−1), resulting from these direct determinations of k1 and k−1 and, also, from the measurements of the equilibrium concentrations of Br, IBr, I, and Br2, is: K1=k1/k−1=161.2±19.7. These data have been used to determine the enthalpy of reaction (1), ΔH298°=−(3.6±0.1) kcal mol−1 and the heat of formation of the IBr molecule, ΔHf,298°(IBr)=(9.8±0.1) kcal mol−1. © 1998 John Wiley & sons, Inc. Int J Chem Kinet 30: 933–940, 1998  相似文献   

3.
Several members of the family of (H2O)5CrR2+ complexes have been shown to react cleanly with the interhalogen molecule IBr. Products of reaction are exclusively the organic iodide RI, Cr(H2O)63+, and Br?-; RBr and (H2O)5CrB2+ are not formed. Rates of the IBr reactions are comparable to those of bromine and much higher than those of iodine. These results are consistent with electrophilic reactions and suggest an SE2 mechanism.  相似文献   

4.
A series of five ternary octanuclear iodine-bromine-chlorine interhalides, [I2Br2Cl4]2− ( 1 ), [I3BrCl4]2− ( 2 ), [I4Br2Cl2]2− ( 3 ), [I2Br4Cl2]2− ( 4 ) and [I3Br3Cl2]2− ( 5 ), have been rationally constructed in two steps. Firstly, addition of a dihalogen (ICl or IBr) to the triaminocyclopropenium chloride salt [C3(NEt2)3]Cl forms the corresponding trihalide salt with [ICl2] or [BrICl] anions, respectively. Secondly, addition of a half-equivalent of a second dihalogen, followed by crystallization at low temperature, gives the corresponding octahalide: addition of Br2 and IBr to [ICl2] gives 1 and 2 , respectively, whereas addition of I2, Br2 and IBr to [BrICl] gives 3 , 4 and 5 , respectively. The five octahalides were characterized by X-ray crystallography and far–IR spectroscopy.  相似文献   

5.
The resonance Raman spectra of gaseous iodine bromide IBr have been studied with the excitation of various argon ion laser lines from 5017 to 4579Å. The fine structures of the fundamental and few overtones of IBr are also studied by various power of 4880Å laser line. The resonance Raman scattering is found to be strong as that of Br2 and ICI. A new term “apparent spectroscopic temperature” is suggusted for the case of the resonance Raman scattering. The apparent spectroscopic temperatures measured in this cell show that the system is not in thermal equilibrium. Br2 is the hottest and I2 is the coldest. IBr is in the middle. Unfortunately, no chemical reaction enhanced phenomenon is found although there should be some chemical reactions occurring under the laser light. The initiating reaction is the photodissociation of the main component IBr which also has large absorptivity. Because of the non-crossing between the B3Π and the 1Π states, the primary products of the photodissociation should be I and Br. The chemical reactions of I and Br with IBr follow. The reactions of I and IBr is endothermic but the reaction of Br with IBr is exothermic. Therefore vibrational hot Br2 is produced and its apparent spectroscopic temperature should be higher. On the other hand, the apparent spectroscopic temperature of I2 is lower.  相似文献   

6.
The nature of halogen bonding is examined via experimental and computational characterizations of a series of associates between electrophilic bromocarbons R? Br (R? Br=CBr3F, CBr3NO2, CBr3COCBr3, CBr3CONH2, CBr3CN, etc.) and bromide anions. The [R? Br, Br?] complexes show intense absorption bands in the 200–350 nm range which follow the same Mulliken correlation as those observed for the charge‐transfer associates of bromide anions with common organic π‐acceptors. For a wide range of the associates, intermolecular R? Br???Br? separations decrease and intramolecular C? Br bond lengths increase proportionally to the Br?→R? Br charge transfer; and the energies of R? Br???Br? bonds are correlated with the linear combination of orbital (charge‐transfer) and electrostatic interactions. On the whole, spectral, structural and thermodynamic characteristics of the [R? Br, Br?] complexes indicate that besides electrostatics, the orbital (charge‐transfer) interactions play a vital role in the R? Br???Br? halogen bonding. This indicates that in addition to controlling the geometries of supramolecular assemblies, halogen bonding leads to electronic coupling between interacting species, and thus affects reactivity of halogenated molecules, as well as conducting and magnetic properties of their solid‐state materials.  相似文献   

7.
The small laser pulse gain method, based on photochemical Br and I lasers, is used to probe 2P3/2 and 2P1/2 states of iodine and bromine atoms in the reactions F + Br2 → BrF + Br (I), I(2P1/2) + Br2 → IBr + Br (II), and Br + IBr → Br2 + I (III). The results obtained are capable of formulating a conservation rule for the spin-orbit excited state.  相似文献   

8.
XPS data of AgBr-coated ion-selective electrodes exposed to high concentrations of Ag+, Cl, Br, I, and NH3 revealed a change in the surface properties of the original electrode. A 40 min to one week exposure of the silver bromide ion-selective electrode surface to solutions containing high concentrations of chloride ions leads to the formation of a mixed halide layer, as the chloride ions are incorporated in the surface. Exposure to high concentrations of iodide-containing solutions results in a new silver iodide layer on top of the original silver bromide laver. Silver ions diffuse to the newly formed layers. NH3 results in the rapid degradation of the AgBr surface as the diamine complex, Ag(NH3)+2, is formed.  相似文献   

9.
The ionic liquid (IL) trihalogen monoanions [N2221][X3] and [N2221][XY2] ([N2221]+=triethylmethylammonium, X=Cl, Br, I, Y=Cl, Br) were investigated electrochemically via temperature dependent conductance and cyclic voltammetry (CV) measurements. The polyhalogen monoanions were measured both as neat salts and as double salts in 1-butyl-1-methyl-pyrrolidinium trifluoromethane-sulfonate ([BMP][OTf], [X3]/[XY2] 0.5 M). Lighter IL trihalogen monoanions displayed higher conductivities than their heavier homologues, with [Cl3] being 1.1 and 3.7 times greater than [Br3] and [I3], respectively. The addition of [BMP][OTf] reduced the conductivity significantly. Within the group of polyhalogen monoanions, the oxidation potential develops in the series [Cl3]>[BrCl2]>[Br3]>[IBr2]>[ICl2]>[I3]. The redox potential of the interhalogen monoanions was found to be primarily determined by the central halogen, I in [ICl2] and [IBr2], and Br in [BrCl2]. Additionally, tetrafluorobromate(III) ([N2221]+[BrF4]) was analyzed via CV in MeCN at 0 °C, yielding a single reversible redox process ([BrF2]/[BrF4]).  相似文献   

10.
Conversion-type batteries apply the principle that more charge transfer is preferable. The underutilized electron transfer mode within two undermines the electrochemical performance of halogen batteries. Here, we realised a three-electron transfer lithium-halogen battery based on I/I+ and Cl/Cl0 couples by using a common commercial electrolyte saturated with Cl anions. The resulting Li||tetrabutylammonium triiodide (TBAI3) cell exhibits three distinct discharging plateaus at 2.97, 3.40, and 3.85 V. Moreover, it has a high capacity of 631 mAh g−1I (265 mAh g−1electrode, based on entire mass loading) and record-high energy density of up to 2013 Wh kg−1I (845 Wh kg−1electrode). To support these findings, experimental characterisations and density functional theory calculations were conducted to elucidate the redox chemistry involved in this novel interhalogen strategy. We believe our paradigm presented here has a foreseeable inspiring effect on other halogen batteries for high-energy-density pursuit.  相似文献   

11.
The addition of bromide ions to the component solutions of the Briggs– Rauscher oscillating system produces a variety of phenomena, depending on the sequence of the addition and on the initial bromide concentration. If the addition is made some minutes after the mixing of H2O2 and acidic IO3 and before adding malonic acid and Mn2+ ions, the oscillations last for five or six cycles, then suddenly ceases. If the addition is made immediately after the mixing of H2O2 and acidic IO3 and before adding malonic acid and Mn2+ ions, the oscillations do not start at all. The addition of bromide ions to an actively oscillating BR reaction causes a rapid suppression of the oscillations. Our observations may be accounted for by a mechanism involving the IBr species. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 641–646, 1998  相似文献   

12.
The kinetics of oxidation of allyl alcohol with potassium bromate in the presence of osmium(VIII) catalyst in aqueous acid medium has been studied under varying conditions. The active species of oxidant and catalyst in the reaction were understood to be Bro3 and H2OsO5, respectively. The autocatalysis exhibited by one of the products, that is, Br, was attributed to complex formation between bromide and osmium(VIII). A composite scheme and rate law were possible. Some reaction constants involved in the mechanism have been evaluated. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 583–589, 1999  相似文献   

13.
《化学:亚洲杂志》2017,12(15):1920-1926
An “in situ sacrifice” process was devised in this work as a room‐temperature, all‐solution processed electrochemical method to synthesize nanostructured NiOx and FeOx directly on current collectors. After electrodepositing NiZn/FeZn bimetallic textures on a copper net, the zinc component is etched and the remnant nickel/iron are evolved into NiOx and FeOx by the “in situ sacrifice” activation we propose. As‐prepared electrodes exhibit high areal capacities of 0.47 mA h cm−2 and 0.32 mA h cm−2, respectively. By integrating NiOx as the cathode, FeOx as the anode, and poly(vinyl alcohol) (PVA)‐KOH gel as the separator/solid‐state electrolyte, the assembled quasi‐solid‐state flexible battery delivers a volumetric capacity of 6.91 mA h cm−3 at 5 mA cm−2, along with a maximum energy density of 7.40 mWh cm−3 under a power density of 0.27 W cm−3 and a maximum tested power density of 3.13 W cm−3 with a 2.17 mW h cm−3 energy density retention. Our room‐temperature synthesis, which only consumes minute electricity, makes it a promising approach for large‐scale production. We also emphasize the in situ sacrifice zinc etching process used in this work as a general strategy for metal‐based nanostructure growth for high‐performance battery materials.  相似文献   

14.
The PIFCO technique in which mass-selected photoion—fluorescence photon coincidences are counted, was used to investigate whether I2+, IBr+ and ICl+ fluoresce. Measurements were made of lifetimes and fluorescence quantum yields of electronic excited states of these ions. Emission was discovered for I2+ and IBr+, but ICl+ apparently does not fluoresce. Information on the radiative properties of Br2+ was obtained as a by-product of the work on IBr+. Fragment ion kinetic energy releases were determined and provide information on dissociative ionization processes in the halogen and interhalogen ions studied.  相似文献   

15.
The title compound, calcium oxide–dialuminium trioxide–calcium dibromide–calcium dichloride hydrate (3/1/0.5/0.5/10), also formulated as Ca2Al(OH)6Br0.478Cl0.522·2H2O (dicalcium aluminium hydro­xide hemibromide hemichloride dihydrate), is a double-layered hydro­xide which belongs to the solid solution Ca2Al(OH)6BrxCl1−x·2H2O, where x can vary from 0 to 1. Chloride and bromide anions of the negatively charged interlayer [Br0.5Cl0.5·2H2O] share statistically the same crystallographic site. Al3+ and Ca2+ cations are coordinated by six and seven O atoms, respectively. All water mol­ecules are bonded to Ca2+ cations and assume the seventh coordination position. Anions in the interlayer are surrounded by ten H atoms. Br and Cl are therefore connected to the main layer by ten hydrogen bonds, six of 2.74 (2) Å and four of 2.52 (5) Å, where the donors are hydroxyl groups and water mol­ecules, respectively. Like the chloride equivalent, the title compound is a 6R polytype with trigonal space group Rc and lattice parameters a = 5.7537 (4) Å and c = 48.108 (4) Å.  相似文献   

16.
In the crystal structure of 2,2′‐bipyridinium(1+) bromide monohydrate, C10H9N2+·Br·H2O, the cation has a cisoid conformation with an intramolecular N—H⋯N hydrogen bond. The cation also forms an N—H⋯O hydrogen bond to an adjacent water mol­ecule, which in turn forms O—H⋯Br hydrogen bonds to adjacent Br anions. In this way, a chain is formed extending along the b axis. Additional interactions (C—H⋯Br and π–π) serve to stabilize the structure further.  相似文献   

17.
Crystals of 1‐(diaminomethylene)thiouron‐1‐ium chloride, C2H7N4S+·Cl, 1‐(diaminomethylene)thiouron‐1‐ium bromide, C2H7N4S+·Br, and 1‐(diaminomethylene)thiouron‐1‐ium iodide, C2H7N4S+·I, are built up from the nonplanar 1‐(diaminomethylene)thiouron‐1‐ium cation and the respective halogenide anion. The conformation of the 1‐(diaminomethylene)thiouron‐1‐ium cation in each case is twisted. Both arms of the cation are planar and rotated in opposite directions around the C—N bonds involving the central N atom. The dihedral angles describing the twisted conformation are 22.9 (1), 15.2 (1) and 4.2 (1)° in the chloride, bromide and iodide salts, respectively. Ionic and extensive hydrogen‐bonding interactions join oppositely charged units into a supramolecular network. The aim of the investigation is to study the influence of the size of the ionic radii of the Cl, Br and I ions on the dimensionality of the hydrogen‐bonding network of the 1‐(diaminomethylene)thiouron‐1‐ium cation. The 1‐(diaminomethylene)thiouron‐1‐ium system should be of use in crystal engineering to form multidimensional networks.  相似文献   

18.
We present the first example of charged imidazolium functionalized porphyrin-based covalent organic framework (Co-iBFBim-COF-X) for electrocatalytic CO2 reduction reaction, where the free anions (e.g., F, Cl, Br, and I) of imidazolium ions nearby the active Co sites can stabilize the key intermediate *COOH and inhibit hydrogen evolution reaction. Thus, Co-iBFBim-COF-X exhibits higher activity than the neutral Co-BFBim-COF, following the trend of F<Cl<Br<I. Particularly, the Co-iBFBim-COF-I showed nearly 100 % CO2 selectivity at a low full-cell voltage of 2.3 V, and achieved a high CO2 partial current density of 52 mA cm−2 with a turnover frequency of 3018 h−1 at 2.4 V in the anion membrane electrode assembly, which is 3.57 times larger than that of neutral Co-BFBim-COF. This work provides new insight into the importance of free anions in the stabilization of intermediates and decreasing the local binding energy of H2O with active moiety to enhance CO2 reduction reaction.  相似文献   

19.
The rate of oxidation of amino acids (AA) by N-Bromoacetamide (NBA) was studied in aqueous buffered medium at 35°C. The rate of disappearance of [NBA] is catalyzed by the Br produced from the reduction of NBA. Analysis of the autocatalyzed reaction gives the kinetic data for the oxidation of bromide ion by NBA. The results suggest that the protonated NBA reacts with Br to form Br2 which rapidly oxidizes amino acids. The rate constant for the reaction between protonated NBA and Br at 35°C is estimated. © 1996 John Wiley & Sons, Inc.  相似文献   

20.
The high theoretical energy density (1274 Wh kg−1) and high safety enable the all-solid-state Na−S batteries with great promise for stationary energy storage system. However, the uncontrollable solid–liquid-solid multiphase conversion and its associated sluggish polysulfides redox kinetics pose a great challenge in tunning the sulfur speciation pathway for practical Na−S electrochemistry. Herein, we propose a new design methodology for matrix featuring separated bi-catalytic sites that control the multi-step polysulfide transformation in tandem and direct quasi-solid reversible sulfur conversion during battery cycling. It is revealed that the N, P heteroatom hotspots are more favorable for catalyzing the long-chain polysulfides reduction, while PtNi nanocrystals manipulate the direct and full Na2S4 to Na2S low-kinetic conversion during discharging. The electrodeposited Na2S on strongly coupled PtNi and N, P-codoped carbon host is extremely electroreactive and can be readily recovered back to S8 without passivation of active species during battery recharging, which delivers a true tandem electrocatalytic quasi-solid sulfur conversion mechanism. Accordingly, stable cycling of the all-solid-state soft-package Na−S pouch cells with an attractive specific capacity of 876 mAh gS−1 and a high energy of 608 Wh kgcathode−1 (172 Wh kg−1, based on the total mass of cathode and anode) at 60 °C are demonstrated.  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