首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The scope of ascorbic acid/iodine and triphenylphosphine/iodine in methanol for the direct reduction of arsenic(V) compounds having the As[dbnd]O group has been investigated. Ascorbic acid/iodine reduces arsonic acids, diphenylarsinic acid (but not dimethylarsinic acid), and triphenylarsine oxide. The rates of reduction depend on the electronic effects of the ligands bound to arsenic and on the hydrogen-bonding strength of the species, when present. When the As(V) compound has an [sbnd]NH 2 or an [sbnd]NH 3 + group, the reduction product reacts with a ketonic form of dehydroascorbic acid, giving condensation product(s). Triphenylphosphine/iodine reduced slowly the zwitterionic o-aminophenylarsonic acid but reduced faster the hydrochloric acid salt of the same acid. It reduced dimethylarsinic acid as well because the powerful electron-withdrawing Ph 3 P +coordinated to As[dbnd]O seems to outweigh the electronic and hydrogen bonding effects.  相似文献   

2.
The direct, nucleophilic imidation of acetanilide derivatives has been performed under mild, iodine(III)‐mediated or ‐catalyzed conditions, employing lithium triflimide as the nitrogen source. The reaction exhibits exclusive regioselectivity for the para position and shows a good tolerance for varied functional groups at both the ortho or meta positions. Preliminary mechanistic data suggest that the LiNTf2 reagent plays a key role in the reactivity.  相似文献   

3.
Summary: Controlled radical polymerization of n‐butyl acrylate by reverse iodine transfer polymerization (RITP) was achieved in ab initio emulsion polymerization to yield a stable and uncolored latex (particle diameter dp = 106 nm). Hydrolysis of iodine, I2, was responsible for an upward deviation from the targeted molecular weight = 10 400 g · mol−1. The iodide concentration [I] was followed by an iodide selective electrode and the amount of efficient iodine (33%) was successfully correlated with the experimental molecular weight = 31 000 g · mol−1. Finally, a simplified mechanism of RITP in ab initio emulsion polymerization taking into account the iodine hydrolysis was proposed.

Evolution of molecular weight and polydispersity index in RITP of BuA in ab initio emulsion.  相似文献   


4.
In reaction of iodine monochloride with CF3COOAg, CH3COONa or (CH3COO)2Pb in acetonitrile and acetic acid the chloride is bonded by metal cations, and electrophilic iodine is generated able to easily iodinate anthracene and carbazole. However at the iodination of anthracene in the presence of oxygen the prevailing process is anthracene oxidation to anthraquinone. In the presence of sulfuric acid protodeiodination of 3-iodocarbazole was found to occur resulting in rearranged products.  相似文献   

5.
A reliable method for direct synthesis of β‐dichlorosubstituted acetanilides is reported. The key transformation involves the oxidative and catalytic cleavage of a carbon‐carbon bond in the presence of iodine trichloride (ICl3). In this protocol ICl3 is used not only as the catalyst but also as the oxidant which widely broadens the scope of its application in organic synthetic chemistry.  相似文献   

6.
M(benzo‐18‐crown‐6)I4 (M = Cd, Hg) are obtained as red columnar crystals from the reactions of benzo‐18‐crown‐6 (b18c6), cadmium and mercury iodide, respectively, and iodine in molar ratios of 1:1:2 in acetonitrile. They both crystallize with the orthorhombic crystal system, P212121, a = 833.7(1), b = 1610.9(1), c = 1846.8(1) pm, V = 2480.3(1) 106·pm3, Z = 4, for M = Cd and a = 823.4(1), b = 1616.5(1), c = 1866.1(1) pm, V = 2483.8(2) 106·pm3 for M = Hg. The crystal structures consist of [M(b18c6)]I2 molecules which are connected to a slightly lengthened iodine molecule via a secondary contact, according to the formulation I2@[MI2@(b18c6)].  相似文献   

7.
It has been shown by electrospray ionization–ion‐trap mass spectrometry that B12I122? converts to an intact B12 cluster as a result of successive stripping of single iodine radicals or ions. Herein, the structure and stability of all intermediate B12In? species (n=11 to 1) determined by means of first‐principles calculations are reported. The initial predominant loss of an iodine radical occurs most probably via the triplet state of B12I122?, and the reaction path for loss of an iodide ion from the singlet state crosses that from the triplet state. Experimentally, the boron clusters resulting from B12I122? through loss of either iodide or iodine occur at the same excitation energy in the ion trap. It is shown that the icosahedral B12 unit commonly observed in dodecaborate compounds is destabilized while losing iodine. The boron framework opens to nonicosahedral structures with five to seven iodine atoms left. The temperature of the ions has a considerable influence on the relative stability near the opening of the clusters. The most stable structures with five to seven iodine atoms are neither planar nor icosahedral.  相似文献   

8.
The liquid-air interface offers a platform for the in-plane growth of free-standing materials. However, it is rarely used for inorganic perovskites and ultrathin non-layered perovskites. Herein the liquid-air interfacial synthesis of inorganic perovskite nanosheets (Cs3Bi2I9, Cs3Sb2I9) is achieved simply by drop-casting the precursor solution with only the addition of iodine. The products are inaccessible without iodine addition. The thickness and lateral size of these nanosheets can be adjusted through the iodine concentration. The high volatility of the iodine spontaneously drives precursors that normally stay in the liquid to the liquid-air interface. The iodine also repairs in situ iodine vacancies during perovskite growth, giving enhanced optical and optoelectronic properties. The liquid-air interfacial growth of ultrathin perovskites provides multi-degree-of-freedom for constructing perovskite-based heterostructures and devices at atomic scale.  相似文献   

9.
The iodine binding capacity (IBC) of glycogen is around 0.30% (w/w) at 3°C. The amount of iodine complexed comprises about 12.5% of the mass of glycogen that takes part in the glycogen–iodine (GI) complex formation. This suggests involvement of four iodine atoms for every 25 anhydroglucose units (AGU, C6H10O5). Since the chromophore is due to the I4 unit within the helix of 11 AGUs, only 44% of the AGUs (11 out of 25) are involved in the complex formation. The heat of formation of the GI complex is around −40 kJ/mol of I2 bonded. These results suggest remarkable similarities with those of the amylopectin–iodine (API) complex. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35: 1409–1412, 1997  相似文献   

10.
A computational study of the mechanism for the iodine(III)‐mediated oxidative amination of alkenes explains the experimentally observed substrate dependence on product distribution. Calculations with the M06 functional have been carried out on the reaction between PhI(N(SO2Me)2)2 and three different representative substrates: styrene, α‐methylstyrene, and (E)‐methylstilbene. All reactions start with electrophilic attack by a cationic PhI(N(SO2Me)2)+ unit on the double bond, and formation of an intermediate with a single C?I bond and a planar sp2 carbocationic center. The major path, leading to 1,2‐diamination, proceeds through a mechanism in which the bissulfonimide initially adds to the alkene through an oxygen atom of one sulfonyl group. This behavior is now corroborated by experimental evidence. An alternative path, leading to an allylic amination product, takes place through deprotonation at an allylic C?H position in the common intermediate. The regioselectivity of this amination depends on the availability of the resonant structures of an alternate carbocationic intermediate. Only in cases where a high electronic delocalization is possible, as in (E)‐methylstilbene, does the allylic amination occur without migration of the double bond.  相似文献   

11.
Plating battery electrodes typically deliver higher specific capacity values than insertion or conversion electrodes because the ion charge carriers represent the sole electrode active mass, and a host electrode is unnecessary. However, reversible plating electrodes are rare for electronically insulating nonmetals. Now, a highly reversible iodine plating cathode is presented that operates on the redox couples of I2/[ZnIx(OH2)4?x]2?x in a water‐in‐salt electrolyte. The iodine plating cathode with the theoretical capacity of 211 mAh g?1 plates on carbon fiber paper as the current collector, delivering a large areal capacity of 4 mAh cm?2. Tunable femtosecond stimulated Raman spectroscopy coupled with DFT calculations elucidate a series of [ZnIx(OH2)4?x]2?x superhalide ions serving as iodide vehicles in the electrolyte, which eliminates most free iodide ions, thus preventing the consequent dissolution of the cathode‐plated iodine as triiodides.  相似文献   

12.
The first comprehensive I(III)―X (X = F, Br, CN, N3, CF3, etc.) bond dissociation energy (BDE) scales for benziodoxol(on)e‐based hypervalent iodine reagents have been developed by virtue of DFT calculations. Excellent correlation is observed between the I(III)―X BDEs and the X―H BDEs, offering a powerful avenue to quickly estimate the group‐transfer ability of a novel benziodoxol(on)e‐based hypervalent reagent.  相似文献   

13.
The reaction of iodine monoxide with chlorine monoxide resulting in atom escape to the gas phase is studied at T = (303 ± 5) K and P = 2.5 Torr using a flow setup for measuring the resonance fluorescence signals of atomic iodine and chlorine. The heterogeneous reaction between chlorine monoxide and iodine monoxide occurring at the reactor surface covered with an F32-L Teflon-like compound and treated by the reaction products is characterized by the rate constant k = (4.9 ± 0.2) × 10–11 cm3 molecule–1 s–1. This value is substantially higher than the rate constant for the homogeneous reaction IO· + ClO· (k 1 1 × 10–12 cm3 molecule–1 s–1).  相似文献   

14.
A series of Zr-based UiO-n MOF materials (n=66, 67, 68) have been studied for iodine capture. Gaseous iodine adsorption was collected kinetically from a home-made set-up allowing the continuous measurement of iodine content trapped within UiO-n compounds, with organic functionalities (−H, −CH3, −Cl, −Br, −(OH)2, −NO2, −NH2, (−NH2)2, −CH2 NH2) by in-situ UV-Vis spectroscopy. This study emphasizes the role of the amino groups attached to the aromatic rings of the ligands connecting the {Zr6O4(OH)4} brick. In particular, the preferential interaction of iodine with lone-pair groups, such as amino functions, has been experimentally observed and is also based on DFT calculations. Indeed, higher iodine contents were systematically measured for amino-functionalized UiO-66 or UiO-67, compared to the pristine material (up to 1211 mg/g for UiO-67-(NH2)2). However, DFT calculations revealed the highest computed interaction energies for alkylamine groups (−CH2NH2) in UiO-67 (−128.5 kJ/mol for the octahedral cavity), and pointed out the influence of this specific functionality compared with that of an aromatic amine. The encapsulation of iodine within the pore system of UiO-n materials and their amino-derivatives has been analyzed by UV-Vis and Raman spectroscopy. We showed that a systematic conversion of molecular iodine (I2) species into anionic I ones, stabilized as I⋅⋅⋅I2 or I3 complexes within the MOF cavities, occurs when I2@UiO-n samples are left in ambient light.  相似文献   

15.
A simple process for the oxy-monofluoromethylation of alkenes is described. In combination with visible-light copper(I) photoredox catalysis, an easily accessible iodine(III) reagent containing monofluoroacetoxy ligands serves as a powerful source of a monofluoromethyl (CH2F) radical, enabling the step economical synthesis of γ-fluoro-acetates from a broad range of olefinic substrates under mild conditions. Applications to late-stage diversification of alkenes derived from complex molecules, amino acids and the synthesis of fluoromethylated heterocycles are also demonstrated.  相似文献   

16.
A new route to synthesize β-iodo amines from sulfonyl aziridines and iodine was developed in the presence of PPh3. This ring-opening reaction was an efficient and simple process to give fl-iodo amines in excellent yields with high chemoselectivity.  相似文献   

17.
Studies on Polyhalides. 40. Ethyltrimethylammonium Pentaiodide (EtMe3N)I5 (EtMe3N)I5 has been prepared by the reaction of stoichiometric amounts of ethyltrimethylammonium iodide and iodine in methanol. It crystallizes in the orthorhombic space group Pna21 with a = 1011.3(1) pm, b = 1255.3(2) pm, c = 1237.7(2) pm and Z = 4. The anionic iodine partial structure is composed of puckered layers which may be derived by deforming a quadratic net with iodide ions in the knots and iodine molecules on the edges of the meshes.  相似文献   

18.
We report the preparation and X‐ray crystallographic characterization of the first crystalline homoatomic polymer chain, which is part of a semiconducting pyrroloperylene–iodine complex. The crystal structure contains infinite polyiodide Iδ?. Interestingly, the structure of iodine within the insoluble, blue starch–iodine complex has long remained elusive, but has been speculated as having infinite chains of iodine. Close similarities in the low‐wavenumber Raman spectra of the title compound and starch–iodine point to such infinite polyiodide chains in the latter as well.  相似文献   

19.
The hypervalent iodine(III) reagent PhI(OTf)2 has been shown to be an effective oxidant for the conversion of linear aryl-hydrazones bearing a pendant alkene into bicyclic diazenium salts. This oxidative cyclization presumably occurs by the iodine(III) mediated formation of a 1-aza-2-azoniaallene salt intermediate that undergoes a subsequent intramolecular 1,3-dipolar cycloaddition with the pendant alkene.  相似文献   

20.
Degradation of the highly reactive hypervalent trifluoromethylation iodine reagent PhI(OAc)(CF3), which can only be generated in situ with mixing PhI(OAc)2 and TMSCF3 in the presence of CsF, was studied by ESI‐MS and GC‐MS combined with 19F‐NMR. The important transient intermediate PhICF3+ was determined by ESI‐MS, and the major volatile products containing CF3 were identified with the authentic compounds by using GC‐MS, such as trifluoromethylbenzene, 2‐iodobenzotrifluoride, 3‐iodobenzotrifluoride, 4‐iodobenzotrifluoride. Meanwhile, more evidences obtained with 19F‐NMR were given for such degradation reaction. A possible rapid CF3 radical transfer reaction pathway was proposed to clarify such degradation progress based on the experimental results. Therefore, this study may be helpful in elucidating the intrinsic reactivity of PhI(OAc)(CF3) and the possible competing side reactions caused by such self‐degradation pathway.  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