首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 390 毫秒
1.

Triaryl trithioarsenites, (ArS)3As, are oxidized by air to As 2 O 3 and ArSSAr. In two cases the parent “thiol” (pyrid-2-thione and 1-hydroxypyrid-2-thione) is coproduced. The oxidation, in nonprotic solvents, is favored by electron-withdrawing groups at the para position of the phenyl group. The products obtained in nonprotic solvents were rationalized by assuming that the binding of the triplet dioxygen to arsenic(III) gives a triplet diradical, (ArS) 3 A?─O─?, or an arsenadioxirane, (ArS) 3 As(O 2 ), intermediate, which decomposes after nucleophilic attack by another (ArS) 3 As molecule. In protic solvents a zwitterion, (ArS) 3 As+─O─O?, and in the presence of moisture a hydroperoxy arsenic(V) compound, (ArS)3As(OH)─O─OH, may be intermediates in the air oxidation of some aromatic trithioarsenites. These data tend to indicate that arsenic(III) bound to suitable groups can directly bind dioxygen, a property which may have implications in chemotherapy and carcinogenesis.  相似文献   

2.

Iodine in dry nonprotic solvents oxidizes triaryl trithioarsenites, (ArS)3As, to As(III) iodide, AsI3, and disulfides, ArSSAr. The reaction most likely involves arylsulfenyl iodide, ArSI, as an intermediate. With triphenyl and tris(4-chlorophenyl) trithioarsenites, AsI3 is prepared in very good yields. When the disulfide, which is produced, has MeCONH─ or ─NH─CMe2─OH groups, then it acts as a Lewis base towards AsI3 forming complexes with stoichiometry 2AsI3·3ArSSAr. Probable coordination modes of the AsI3 are discussed.  相似文献   

3.
Octasulfur (elemental sulfur) does not react at room temperature with a variety of AsIII compounds of the type Ar‐As(SPh)2 and R‐As(SPh)2. However, in the presence of catalytic amounts of triethylamine, which probably acts as an activator of octasulfur, reactions do take place. The products isolated from the aromatic dithioarsonites are not the expected Ar‐As(S)(SPh)2 but the sulfide, cyclo‐(PhAsS)4, and the sesquisulfides, (ArAs)2S3, which are the same with those obtained by reduction of arylarsonic acids with hydrogen sulfide. The action of S8/Et3N on aliphatic dithioarsonites did not give any AsV products but gave mixtures of non‐separable oligomeric (RAsS)x. Probable mechanistic routes for these reactions are proposed. The known cyclo‐(PhAsS)4 and (PhAs)2S3 and the new (2‐O2N‐C6H4‐As)2S3 have been structurally characterized: (2‐O2N‐C6H4‐As)2S3, monoclinic, C2/c, a = 13.564(9)Å, b = 7.982(6)Å, c = 16.67(1)Å, β = 117.63(2)°, V = 1599(2)Å3, Z = 4, wR2 0.0640. The 1, 4‐diarsa‐2, 3, 5‐trithiacyclopentane ring is puckered and the two AsIII atoms are pseudo tetrahedral.  相似文献   

4.
This study is to prepare a nanosuspension based on poly(lactic‐co‐glycolic acid) (PLGA) for delivery, controlled release and enhanced anti‐solid tumor effects of As2O3. As2O3‐loaded PLGA nanocapsules (As2O3‐PLGA NCs) were prepared by double emulsion‐solvent evaporation method and were optimized by univariate analysis in combination with orthogonal experimental according to several factors. The optimized As2O3‐PLGA NCs presented suitable physical stability, favorable size of (200.2±10.6) nm (PDI=0.117±0.008), spherical shape, and high encapsulation efficiency (92.48%±2.14%). The in vitro suspension stability of the NCs was excellent. The release of As2O3 from the NCs showed pH responsive release characteristics. The NCs can be efficiently taken up by SMMC‐7721 cell and showed excellent antitumor efficacy against SMMC‐7721 cell line. Then, As2O3‐PLGA NCs could be considered as a promising formulation for the pH dependent release of As2O3 in cancer cells and enhance the anti‐solid tumor effects of As2O3.  相似文献   

5.
Bunsen's cacodyl disulfide, Me2As(S)‐S‐AsMe2 ( 1 ), reacted with iodine giving the novel dimethylarsinosulfenyl iodide, Me2As‐S‐I ( 3 ) although theoretical calculations indicated that the AsV compound Me2As(S)‐I ( 4 ) was more stable in the gas phase. The oily product was stable neat and as a solution in CDCl3 at +4 °C and –20 °C for at least 15 d. Light, H2O, H2O2, and Zn dust, but not NaI or Ag, decomposed it. Compound 3 did not interact with Ph3N, with Ph2NH and PhNH2 it interacted but not reacted. 3 was decomposed by piperidine, with pyridine and 4‐dimethylaminopyridine it interacted and produced Me2As‐SS‐AsMe2 ( 2 ) and I2 that formed charge transfer complexes Base · I2, whereas Et3N decomposed 3 , and 3Et3N · 2I2 was isolated. 3 was desulfurized by Ph3P and (Me2N)3P completely, and by (PhO)3P and (PhS)3P partially. The reactions of 3 with (Me2N)3P, (PhS)3P, and (EtO)3P were complicated. From the AsIII nucleophiles, only Ph3As was bound, while (PhS)3As reacted slowly in a complicated manner with 3 . No interaction of 3 with MeOH or PhOH was observed but NaOH, Ag2O, and PhONa decomposed it. Thiophenol produced traces of Me2As‐SPh ( 10 ) and sodium thiophenolate attacked mainly at AsIII of 3 . Thus, externally stabilized sulfenium ions of the type Me2As‐S‐Nu+I were not obtained.  相似文献   

6.
Two isostructural diarsenates, SrZnAs2O7 (strontium zinc diarsenate), (I), and BaCuAs2O7 [barium copper(II) diarsenate], (II), have been synthesized under hydrothermal conditions and characterized by single‐crystal X‐ray diffraction. The three‐dimensional open‐framework crystal structure consists of corner‐sharing M2O5 (M2 = Zn or Cu) square pyramids and diarsenate (As2O7) groups. Each As2O7 group shares its five corners with five different M2O5 square pyramids. The resulting framework delimits two types of tunnels aligned parallel to the [010] and [100] directions where the large divalent nine‐coordinated M1 (M1 = Sr or Ba) cations are located. The geometrical characteristics of the M1O9, M2O5 and As2O7 groups of known isostructural diarsenates, adopting the general formula M1IIM2IIAs2O7 (M1II = Sr, Ba, Pb; M2II = Mg, Co, Cu, Zn) and crystallizing in the space group P21/n, are presented and discussed.  相似文献   

7.
The syntheses and X‐ray single‐crystal low‐temperature structures of the four new ammoniates [Li(NH3)4]3As7?NH3 ( 1 ), [Rb(18‐crown‐6)]3As7?8 NH3 ( 2 ), Cs3As7?6 NH3 ( 3 ), and (Ph4P)2CsAs7?5 NH3 ( 4 ) are reported. The compounds were obtained by either direct reduction of As with Li/Cs in liquid NH3, solvation of Cs4As6/Rb4As6 in liquid NH3, or by extraction of solid Cs3As7. While compound 1 contains isolated As polyanions, As? M contacts (M=Na?Cs) lead to neutral [Rb(18‐crown‐6)]3As7 units in 2 , a three‐dimensional, extended network in 3 , and one‐dimensional, infinite [CsAs7]2? chains in 4 , respectively.  相似文献   

8.
An inorganic–organic hybrid compound, (H2bpy)3[AsIIIAs2 VMo15 VIMo3 VO62]·3H2O (bpy: 4,4′-bipyridine) (1) has been synthesized under hydrothermal conditions and characterized by elemental analysis, X-ray photoelectron spectroscopy (XPS), X-ray powder diffraction, IR spectrum, UV–Vis spectrum, thermogravimetric analysis and single-crystal X-ray diffraction. The crystallographic analysis reveals that compound 1 is composed of a Wells–Dawson polyoxoanion [As2 VMo15 VIMo3 VO62]9? with the mixed valence of MoV,VI, which acts as a tetradentate ligand coordinating with the AsIII cation to form the mixed valence AsIII,V containing arsenomolybdate with the single cap structure in a chelate coordination mode. In the solid state, compound 1 shows a 3D supramolecular structure through hydrogen-bonding interactions (C–H···O, N–H···O, N–H···N). In addition, compound 1 exhibits reversible multi-electron redox processes and effective electrocatalytic activities towards the reduction of H2O2, NaNO2 and KBrO3. Moreover, compound 1 is used as a reducing agent of the graphene oxide to prepare graphene by a green chemistry type one-step synthesis method, which is characterized by XPS, Raman spectroscopy, PXRD, IR, scanning electron microscopy and transmission electron microscopy.  相似文献   

9.
The Chlorooxoarsenates(III) (PPh4)2[As4O2Cl10] · 2 CH3CN and (PPh4)2[As2OCl6] · 3 CH3CN (PPh4)2[As2Cl8] can be prepared from As2O3, SOCl2 and PPh4Cl in acetonitrile. Its oxidation with chlorine yields PPh4[AsCl6]. This was also obtained directly from arsenic, chlorine and PPh4Cl, (PPh4)2[As4O2Cl10] · 2 CH3CN being a side product; the latter was obtained with high yield from AsCl3, As2O3 and PPh4Cl in acetonitrile. By addition of PPh4Cl it was converted to (PPh4)2[As2OCl6] · 3 CH3CN. According to their X-ray crystal structure analyses, both crystallize in the triclinic space group P 1. The [As4O2Cl10]2– ion can be regarded as a centrosymmetric association product of two Cl2AsOAsCl2 molecules and two Cl ions, each Cl ion being coordinated with all four As atoms. In the [As2OCl6]2– ion the As atoms are linked via the O atom and two Cl atoms.  相似文献   

10.
A family of three sandwich‐type, phenylantimony(III)‐containing tungstoarsenates(III), [(PhSbIII){Na(H2O)}AsIII2W19O67(H2O)]11? ( 1 ), [(PhSbIII)2AsIII2W19O67(H2O)]10? ( 2 ), and [(PhSbIII)3(B‐α‐AsIIIW9O33)2]12? ( 3 ), have been synthesized by one‐pot procedures and isolated as hydrated alkali metal salts, Cs3K3.5Na4.5[(PhSbIII){Na(H2O)}AsIII2W19O67(H2O)]?41H2O ( CsKNa ‐ 1 ), Cs4.5K5.5[(PhSbIII)2AsIII2W19O67(H2O)]?35H2O ( CsK‐2 ), and Cs4.5Na7.5[(PhSbIII)3(B‐α‐AsIIIW9O33)2]?42H2O ( CsNa ‐ 3 ). The number of incorporated {PhSbIII} units could be selectively tuned from one to three by careful control of the reaction parameters. The three compounds were characterized in the solid state by single‐crystal XRD, IR spectroscopy, and thermogravimetric analysis. The aqueous solution stability of sandwich polyanions 1 – 3 was also studied by multinuclear (1H, 13C, 183W) NMR spectroscopy. Effective inhibitory activity against six different kinds of bacteria was identified for all three polyanions, for which the activity increased with the number of incorporated {PhSbIII} groups.  相似文献   

11.
Nanostructured Sr2As2O7 samples were synthesized by hydrothermal crystal growth reactions between Sr(NO3)2 and As2O3 with different stoichiometric Sr:As molar ratios as 1:1 (S1), 1:2 (S2), and 2:1 (S3). The synthesized nanomaterials were characterized by powder X‐ray diffraction (PXRD) technique and fourier‐transform infrared (FT‐IR) spectroscopy. Rietveld analysis showed that the obtained materials were crystallized well in the monoclinic crystal structure with the space group P21. The morphologies of the synthesized materials were studied by field emission scanning electron microscopy (FESEM) technique. TEM images verified formation of the particles with nanometer size. Ultraviolet‐visible spectra analysis showed that the synthesized Sr2As2O7 nanomaterials had strong light absorption in the ultraviolet light region. Photocatalytic performance of the synthesized nanomaterials was also investigated for the degradation of pollutant Malachite Green (4‐{[4‐(dimethylamino)phenyl](phenyl)methylidene}‐N′N‐dimethylcyclohexa‐25‐dien‐1‐iminium chloride) (MG) in aqueous solution under solar light condition. The optimal conditions were obtained by design expert software for S1. It was found that the optimum condition was 0.14 mL H2O2, 20 mg catalyst, and 33 min time. The volume and concentration of the as prepared MG solution were 70 mL and 100 ppm, respectively, for obtaining the optimum conditions. The degradation yield in the optimized conditions was 97 % for S1.  相似文献   

12.
The first members of a promising new family of hybrid amino acid–polyoxometalates have emerged from a search for modular functional molecules. Incorporation of glycine (Gly) or norleucine (Nle) ligands into an yttrium‐tungstoarsenate structural backbone, followed by crystallization with p‐methylbenzylammonium (p‐MeBzNH3+) cations, affords (p‐MeBzNH3)6K2(GlyH)[AsIII4(YIIIWVI3)WVI44YIII4O159(Gly)8‐ (H2O)14] ? 47 H2O ( 1 ) and enantiomorphs (p‐MeBzNH3)15(NleH)3 [AsIII4(MoV2MoVI2)WVI44YIII4O160(Nle)9(H2O)11][AsIII4(MoVI2WVI2)‐ WVI44YIII4O160(Nle)9(H2O)11] (generically designated 2 : L ‐Nle, 2 a ; D ‐Nle, 2 b ). An intensive structural, spectroscopic, electrochemical, magnetochemical and theoretical investigation has allowed the elucidation of site‐selective metal substitution and photoreduction of the tetranuclear core of the hybrid polyanions. In the solid state, markedly different crystal packing is evident for the compounds, which indicates the role of noncovalent interactions involving the amino acid ligands. In solution, mass spectrometric and small‐angle X‐ray scattering studies confirm maintenance of the structure of the polyanions of 2 , while circular dichroism demonstrates that the chirality is also maintained. The combination of all of these features in a single modular family emphasizes the potential of such hybrid polyoxometalates to provide nanoscale molecular materials with tunable properties.  相似文献   

13.
R3SnSAr are oxidised by H2O2, 2,4,4,6-tetrabromocyclohexa-2,5-dienone and NaIO4 (in 1 : 1 ratio) to disulphides, ArSSAr. With NaIO4 as the oxidant, Ph3SnSAr, gives a complex triphenyltin periodate; however, with the other two oxidants, (Ph3Sn)2O is produced. Ph3GeSAr was oxidised by NaIO4 to (Ph3Ge)2O and ArSSAr.Triorgano-germanium and -tin O-sulphinates, R3MOSOAr react with arenesulphenyl chlorides, ArSCl, giving thiosulphonates, ArSS(O2)Ar, and with sulphur dichloride, producing di(sulphonyl) sulphides ArS(O2)SS(O2)Ar. Triorgano-germanium p-toluene sulphonate readily formed by oxidation of the O-sulphinate derivative, does not react with arenesulphenyl chlorides.  相似文献   

14.
《Electroanalysis》2003,15(7):613-619
The interaction of arsenic trioxide (As2O3) with calf thymus double‐stranded DNA (dsDNA), calf thymus single‐stranded DNA (ssDNA) and also 17‐mer short oligonucleotide (Probe A) was studied electrochemically by using differential pulse voltammetry (DPV) with carbon paste electrode (CPE) at the surface and also in solution. Potentiometric stripping analysis (PSA) was employed to monitor the interaction of As2O3 with dsDNA in solution phase by using a renewable pencil graphite electrode (PGE). The changes in the experimental parameters such as the concentration of As2O3, and the accumulation time of As2O3 were studied by using DPV; in addition, the reproducibility data for the interaction between DNA and As2O3 was determined by using both electrochemical techniques. After the interaction of As2O3 with dsDNA, the DPV signal of guanine was found to be decreasing when the accumulation time and the concentration of As2O3 were increased. Similar DPV results were also found with ssDNA and oligonucleotide. PSA results observed at a low DNA concentration such as 1 ppm and a different working electrode such as PGE showed that there could be damage to guanine bases. The partition coefficients of As2O3 after interaction with dsDNA and ssDNA in solution by using CPE were calculated. Similarly, the partition coefficients (PC) of As2O3 after interaction with dsDNA in solution was also calculated by PSA at PGE. The features of this proposed method for the detection of DNA damage by As2O3 are discussed and compared with those methods previously reported for the other type of DNA targeted agents in the literature.  相似文献   

15.
Two novel As‐V‐O cluster supported transition metal complexes, [Zn(en)2][Zn(en)2(H2O)2][{Zn(en)(enMe)}As6V15O42(H2O)]·4H2O ( 1 ) and [Zn2(enMe)2(en)3][{Zn(enMe)2}As6V15O42(H2O)]·4H2O ( 2 ), have been hydrothermally synthesized. The single X‐ray diffraction studies reveal that both compounds consist of discrete noncentral polyoxoanions [{Zn(en)(enMe)}As6V15O42(H2O)]4? or [{Zn(enMe)2}As6V15O42(H2O)]4? cocrystallized with respective zinc coordination complexes. Interestingly, compounds 1 and 2 exhibit the first two polyoxovanadates containing As8V15O42‐(H2O)]6? cluster decorated by only one transition metal complex. Crystal data: 1 , monoclinic, P21/n, a = 14.9037(4) Å, b = 18.1243(5) Å, c = 27.6103(7) Å, β = 105.376(6)°, Z = 4; 2 monoclinic, P21/n, a = 14.9786(7) Å, b = 33.0534(16) Å, c = 14.9811(5) Å, Z = 4.  相似文献   

16.
Concerning Ba2As6O11 Ba2As6O11 was prepared by hydrothermal reaction of BaO with As2O3 at a temperature of 200°C. An X-ray structural analysis demonstrated that the phase contains highly condensed polyarsenate(III) anions (As6O114?)n. A zweier double chain structure was observed for the anion, in which the individual chains are bridged to one another by two Ψ-AsO3 tetrahedra, so that As10O10 rings are formed. The double chains are linked into “double sheets”, perpendicular to the b-axis, through relatively strong secondary O … As bonds with an average length of 2.68 Å.  相似文献   

17.
Polycationic Hg–As Frameworks with Trapped Anions. II Synthesis, Crystal Structure, and Magnetism of (Hg6As4)[MoCl6]Cl, (Hg6As4)[TiCl6]Cl, and (Hg6As4)[TiBr6]Br (Hg6As4)[MoCl6]Cl is obtained by reaction of Hg2Cl2, Hg, As, and MoCl4 in closed, evacuated glass ampoules in a temperature gradient 450 → 400 °C in form of dark red cubelike crystals. (Hg6As4)[TiCl6]Cl and (Hg6As4)[TiBr6]Br are also formed in closed, evacuated ampoules from Hg2X2 (X = Cl, Br), Hg, As, and Ti metal at 275 °C and 245 °C in form of dark green and black crystals, respectively. All three compounds are air and light sensitive. They crystallize isotypically (cubic, Pa 3, a = 1207.8(4) pm for (Hg6As4)[MoCl6]Cl, a = 1209.4(3) pm for (Hg6As4)[TiCl6]Cl, a = 1230.9(3) pm for (Hg6As4)[TiBr6]Br, Z = 4). The structures consist of a three‐dimensionally connected Hg–As framework which is made up of As2 groups (As–As distance averaged 242 pm) each connected via six Hg atoms to six neighbouring As2 groups. There are two cavities of different size in the polycationic framework. The bigger cavity is filled with [MoCl6]3–, [TiCl6]3–, and [TiBr6]3– ions of nearly ideal octahedral shape, the smaller cavity with discrete halide ions. The magnetic properties of the two Ti containing compounds are in accordance with a d1 paramagnetism. The temperature dependence and the magnitude of the magnetic moment can be interpreted with consideration of the spin‐orbit coupling. The so far known representatives of this structure type can be characterised by the ionic formula (Hg6Y4)4+[MX6]3–X (Y = As, Sb; M = Sb3+, Bi3+, Mo3+, Ti3+; X = Cl, Br).  相似文献   

18.
Sm2As4O9: An Unusual Samarium(III) Oxoarsenate(III) According to Sm4[As2O5]2[As4O8] Pale yellow single crystals of the new samarium(III) oxoarsenate(III) with the composition Sm4As8O18 were obtained by a typical solid‐state reaction between Sm2O3 and As2O3 using CsCl and SmCl3 as fluxing agents. The compound crystallizes in the triclinic crystal system with the space group (No. 2, Z = 2; a = 681.12(5), b = 757.59(6), c = 953.97(8) pm, α = 96.623(7), β = 103.751(7), γ = 104.400(7)°). The crystal structure of samarium(III) oxoarsenate(III) with the formula type Sm4[As2O5]2[As4O8] (≡ 2 × Sm2As4O9) contains two crystallographically different Sm3+ cations, where (Sm1)3+ is coordinated by eight, but (Sm2)3+ by nine oxygen atoms. Two different discrete oxoarsenate(III) anions are present in the crystal structure, namely [As2O5]4? and [As4O8]4?. The [As2O5]4? anion is built up of two Ψ1‐tetrahedra [AsO3]3? with a common corner, whereas the [As4O8]4? anion consists of four Ψ1‐tetrahedra with ring‐shaped vertex‐connected [AsO3]3? pyramids. Thus at all four crystallographically different As3+ cations stereochemically active non‐binding electron pairs (“lone pairs”) are observed. These “lone pairs” direct towards the center of empty channels running parallel to [010] in the overall structure, where these “empty channels” being formed by the linkage of layers with the ecliptically conformed [As2O5]4? anions and the stair‐like shaped [As4O8]4? rings via common oxygen atoms (O1 – O6, O8 and O9). The oxygen‐atom type O7, however, belongs only to the cyclo‐[As4O8]4? unit as one of the two different corner‐sharing oxygen atoms.  相似文献   

19.
Polynuclear Iron/Tantalum and Tantalum Complexes with Asn Ligands Starting with [Cp@Ta(CO)4] ( 1 ) (Cp@ = C5H3tBu2‐1,3) and As4 or (tBuAs)4 ( 2 ) its thermolysis at 190 °C in decalin gives [{Cp@Ta}2(μ‐η4 : η4‐As8)] ( 3 ), which is also formed according to equation (2) in addition to [{Cp@Ta}3As6] ( 5 ). The reaction of 1 or [{Cp*(OC)2Fe}2] ( 6 ) with 3 affords 5 or [{Cp*Fe}{Cp@Ta}As5] ( 8 ) demonstrating the use of 3 as Asn source. 8 can also be synthesized from 1 and [Cp*Fe(η5‐As5)] ( 7 ) for which the cothermolysis of 2 and 6 gives a better yield.  相似文献   

20.
Gas‐phase reactions of model carbosulfonium ions (CH3‐S+ = CH2; CH3CH2‐S+ = CH2 and Ph‐S+ = CH2) and an O‐analogue carboxonium ion (CH3‐O+ = CH2) with acyclic (isoprene, 1,3‐butadiene, methyl vinyl ketone) and cyclic (1,3‐cyclohexadiene, thiophene, furan) conjugated dienes were systematically investigated by pentaquadrupole mass spectrometry. As corroborated by B3LYP/6‐311 G(d,p) calculations, the carbosulfonium ions first react at large extents with the dienes forming adducts via simple addition. The nascent adducts, depending on their stability and internal energy, react further via two competitive channels: (1) in reactions with acyclic dienes via cyclization that yields formally [4 + 2+] cycloadducts, or (2) in reactions with the cyclic dienes via dissociation by HSR loss that yields methylenation (net CH+ transfer) products. In great contrast to its S‐analogues, CH3‐O+ = CH2 (as well as C2H5‐O+ = CH2 and Ph‐O+ = CH2 in reactions with isoprene) forms little or no adduct and proton transfer is the dominant reaction channel. Isomerization to more acidic protonated aldehydes in the course of reaction seems to be the most plausible cause of the contrasting reactivity of carboxonium ions. The CH2 = CH‐O+ = CH2 ion forms an abundant [4 + 2+] cycloadduct with isoprene, but similar to the behavior of such α,β‐unsaturated carboxonium ions in solution, seems to occur across the C = C bond. Copyright © 2012 John Wiley & Sons, Ltd.  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-23

京公网安备 11010802026262号