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 0 毫秒
1.
The molecular structure of the title complexes [Fe(H2O)4][Fe(Hedta)(H2O)]2 · 4H2O (I) and [Fe(H[2edta)(H2O)] · 2H2O (II) have been determined by single-crystal X-ray analyses. The crystal data are as follows: I: monoclinic, P21/n, A = 11.794(2), B = 15.990(2), C = 9.206(2) Å, β = 90.33(1)°, V = 1736.1(5) Å3, Z = 2 and R = 0.030; II: monoclinic, C2/c, A = 11.074(2), B = 9.856(2), C = 14.399(2) Å, β = 95.86(1)°, V = 1563.3(4) Å3, Z = 4 and R = 0.025. I is found to be isomorphous with the MnII analog reported earlier and to contain a seven-coordinate and approximately pentagonal-bipyramidal (PB) [FeII(Hedta)(H2O] unit in which Hedta acts as a hexadentate ligand. The [FeII(H2edta)(H2O)] unit in II has also a seven-coordinate PB structure with the two protonated equatorial glycine arms both remaining coordinated, and thus bears a structural resemblance to the seven-coordinate [CoII(H2edta)(H2O)] reported previously.  相似文献   

2.
The chlorocadmate(II) systems of (H2me2pipz)[Cd2Cl6(H2O)2] (1) and (H2mepipz)2[Cd3Cl10(H2O)] (2) (L = me2pipz = N,N′-dimethylpiperazine; L′ = mepipz = N-methylpiperazine) were prepared and their structural and thermal properties investigated. Compound 1 is monoclinic, space group P21/c, A = 7.664(1), B = 7.472(4), C = 15.347(1) Å, β = 99.468(7)°, Z = 2, R = 0.024. The crystal structure consists of organic cations and infinite one-dimensional chains of [CdCl3(H2O)]n3− anions. Each Cd atom is octahedrally surrounded by bridged and terminal chlorine atoms and by a water molecule, which is in trans position with respect to the terminal chlorine atom. Inter- and intrachain hydrogen bond interactions between the terminal chlorine atoms and the water molecules contribute to the crystal packing. Compound 2 is orthorhombic, space group Cmc21, A = 15.286(3), B = 13.354(3), C = 13.154(3) Å, R = 0.023. The crystal structure consists of organic dications and infinite chains of [Cd2Cl6(CdCl4H2O]n4− units running along the [001] axis. Each unit is formed of regularly alternate six-coordinated Cd atoms, one of them linking one pentacoordinated Cd atom which completes its coordination througha water molecule. A strong hydrogen bond interaction involving the organic dication and the inorganic chain contributes to the crystal packing. Differential hydrogen bond interaction involving the organic dication and the inorganic chain contributes to the crystal packing. Differential scanning calorimetry measurements did not show the presence of any structural phase transitions. The structures are compared with those of (H2pipz)[Cd2Cl6(H2O)2] (3), (H2mepipz)[Cd2Cl6(H2O)2]·H2O (4) and (H2mepipz)[Cd2Cl6] (5) (L = pipz = piperazine, L′ = mepipz = N-ethylpiperazine).  相似文献   

3.
The interaction between Ac-AMP2, a lectin-like small protein with antimicrobial and antifungal activity isolated from Amaranthus caudatus, and N,N′,N″-triacetyl chitotriose was studied using 1H NMR spectroscopy. Changes in chemical shift and line width upon increasing concentration of N,N′,N″-triacetyl chitotriose to Ac-AMP2 solutions at pH 6.9 and 2.4 were used to determine the interaction site and the association constant Ka. The most pronounced shifts occur mainly in the C-terminal half of the sequence. They involve the aromatic residues Phe18, Tyr20 and Tyr27 together with their surrounding residues, as well as the N-terminal Val-Gly-Glu segment. Several NOEs between Ac-AMP2 and the N,N′,N″-triacetyl chitotriose resonances are reported.  相似文献   

4.
[MnL](ClO4)2 (L = N,N′,N″-tris(2-hydroxypropyl)-1,4,7-triazacyclononane) has been tested for catalyzing sulfide oxidation. In the presence of this complex, ethyl phenyl sulfide, butyl sulfide and phenyl sulfide are completely oxidized to the corresponding sulfoxides and sulfones with H2O2 as the oxidant. 2-Chloroethyl phenyl sulfide oxidation yield 2-chloroethyl phenyl sulfone and phenyl vinyl sulfone. In ethyl phenyl sulfide oxidation, effects of complex and H2O2 concentration and temperature on the reaction rate have been discussed. Through controlling reaction conditions, ethyl phenyl sulfoxide and ethyl phenyl sulfone may be produced selectively. The UV–Vis and electron paramagnetic resonance (EPR) studies on catalyst solution indicate that metal centre of the complex is transformed from Mn(II) to Mn(IV) after the addition of H2O2. At 25 °C, rate constant for ethyl phenyl sulfide oxidation is 4.38 × 10−3 min−1.  相似文献   

5.
The ligand N, N′-bis[2,2-dimethyl-4-(2-hydroxyphenyl)-3-aza-3-buten] oxamide with two identical coordination sites reacts with copper ions in its tetradeprotonated form to yield the dinuclear complex [Cu2(C24H26N4O4)]·H2O. The structure of this compound has been determined by the X-ray diffraction method. The crystals are orthorhombic with a = 11.744(1), B = 16.369(2), C = 26.340(3) Å, V = 5064(1) Å3, Z = 8, space group Pbca. The oxamide is in a trans conformation with two different environments for the copper centres, a (4 + 1) coordination mode for the first one and a square planar environment for the other one. The water molecule is not directly bound to a copper centre, but involved in hydrogen bonding with the two oxygen atoms of an N2O2 coordination site. Indeed, extra coordination comes from a phenolic oxygen atom belonging to an adjacent dinuclear unit. Static susceptibility measurements point to a strong intrapair antiferromagnetic exchange interaction of 2J = −520(±4) cm−1 and possibly an interpair ferromagnetic exchange interaction of 10(±5) cm−1.  相似文献   

6.
The X-ray structure is reported for the complex Cu2(medpco-2H)Cl2, (medpco = N,N′-bis-N,N-dimethylaminoethyl)pyridine-2,6-dicarboxamide 1-oxide. The complex is triclinic, , a=8.313(4), B=11.403(5), C=11.611(3) Å, =91.66(3), β=108.99(4), γ=109.60(3)° and Z=2. The deprotonated ligand (medpco-2H)2− acts as a binulceating ligand, producing an N-oxide-bridged complex. Each copper in Cu2(medpco-2H)Cl2 is five-coordinate, being coordinated by a bridging N-oxide oxygen, a deprotonated amide nitrogen, a tertiary amine nitrogen and two bridging chlorides. The complex does not exhibit significant magnetic interaction, and this may be the result of distortion of the bridging geometry from planarity. A range of other, apparently N-oxide-bridged, complexes of the type Cu2(medpco-2H)X2 is reported. The complex Cu2(medpco-2H)Br2·H2O is strongly antiferromagnetic, with magnetic data closely fitting the expected binuclear structure.  相似文献   

7.
A new monohelical OH bridged dinuclear complex [Zn2(dmqpy)(OOCCH3)2(μ-OH)][ClO4] · 0.5EtOH, where dmqpy is 6,6-dimethyl-2,2′:6′,2″:6″,2:6,2-quinquepyridine, has been synthesized and characterized by X-ray crystallography: monoclinic, space group P21/c, a=13.670(1), b=14.751(1), c=16.782(1) Å, β=96.59(1)°, U=3361.7(4) Å3, Z=4, R=0.0601. Two Zn(II) ions are in different coordination modes, one is five-coordinate with a N3O2 donor set and the other is N2O2 four-coordinate with a distorted tetrahedral geometry, and the zinc ions are bridged by a hydroxyl group. The presence of the OH bridge is further confirmed by electrospray mass and infrared spectroscopies. The solution properties of the complex were investigated by 1H NMR spectroscopy. The results of NMR indicate that the complex has higher symmetry in solution than in the solid state.  相似文献   

8.
We have analyzed the effect of N,N′-bis-(2,3-methylenedioxyphenyl)urea (2,3-MDPU) and N,N′-bis-(3,4-methylenedioxyphenyl)urea (3,4-MDPU), two symmetrically substituted diphenylurea derivatives with no auxin or cytokinin-like activity, on the rooting capacity of Pinus radiata stem cuttings. Results indicate that both diphenylurea derivatives enhance adventitious rooting in the presence of exogenous auxin (indole-3-butyric acid, IBA), even at low auxin concentration, in rooting-competent cuttings, but have no effect on the adventitious rooting of low or null competent-to-root cuttings. Histological analyses show that, in the simultaneous presence of MDPUs and low concentration of exogenous auxin, adventitious root formation is induced in the cell types that retain intrinsic competence to form adventitious roots in response to auxin. The time course of cellular events leading to root formation and the time of root emergence are closely similar to that observed in cuttings treated only with higher auxin concentration. In addition, the mRNA level of a P. radiata SCARECROW-LIKE gene, which is significantly induced in the presence of the optimal concentration (10 μM) of exogenous auxin needed for cuttings to root, is increased in the presence of MDPUs and low concentration of exogenous auxin (1 μM). The expression of a P. radiata SHORT-ROOT gene in rooting-competent cuttings during adventitious rooting is also affected by the presence of MDPUs when combined with auxin. As MDPUs do not affect the expression of either gene in the absence of exogenous auxin, but only in its presence, we suggest that MDPUs could interact, directly or indirectly, with the auxin-signalling pathways in rooting-competent cuttings during adventitious rooting.  相似文献   

9.
The stepwise synthesis of mononuclear (4f) and heterodinuclear (3d–4f) Salen-like complexes has been investigated through structural determination of the intermediate and final products occurring in the process. In the first step, reactions of ligand H2L and Ln(NO3)3 · 6H2O give rise to three mononuclear lanthanide complexes Ln(H2L)(NO3)3 [H2L = N,N′-ethylene-bis(3-methoxysalicylideneimine), Ln = Nd (1), Eu (2) and Tb (3)], in which N,N′-ethylene-bis(3-methoxysalicylideneimine) acts as tetradentate ligands with the O2O2 set of donor atoms capable of effective coordination. These species are fairly stable and have been isolated. Then, addition of Cu(Ac)2 · H2O to the mononuclear lanthanide complex yields expected heterodinuclear (3d–4f) complexes Cu(L)Ln(NO3)3 · H2O [Ln = Nd (4) and Eu (5)] where the Cu(II) ion is inserted to the inner N2O2 cavity. Luminescent analysis reveals that complex 3 exhibits characteristic metal-centered fluorescence of Tb(III) ion. However, the characteristic luminescence of both Sm(III) and Eu(III) ions is not observed both in solution and solid state of the complexes.  相似文献   

10.
11.
The ligand 1,2-bis(3,5-dimethylpyrazol-1-yl)ethane has been synthesized by the direct reaction of 1,2-dibromoethane and 3,5-dimethylpyrazole. The complex of this ligand with palladium(II) chloride has been prepared and the structure of its toluene solvate has been determined by X-ray crystallography. The compound crystallises in the triclinic space group , a = 11.088(5), B = 13.786(5), C = 21.169(9) Å, = 86.96(3), β = 81.02(3), γ = 73.32(3)° with Z = 2. Final residuals after least-squares refinement: R = 0.030, Rw = 0.039. The compound has a trimeric structure which may be described as a ‘molecular tricorn’: the ligand bridges adjacent palladium centres, giving rise to a 21-membered trimetallic macrocycle. The overall structure closely approximates D3 symmetry with approximate two-fold axes passing through each palladium atom and the centre of the 1,2-ethanediyl moiety opposite. An interesting feature of the structure is the close approach of several hydrogen atoms from the 1,2-ethanediyl groups to each palladium centre; these interactions are thought to be close-packing rather than agostic bonds.  相似文献   

12.
Diorganozinc compounds R2Zn (R=alkyl or aryl) react with N,N′-bis(2,6-di-isopropylphenyl)-1,4-diaza-1,3-butadiene, (i-Pr2Ph)N=CHCHp=N(i-Pr2Ph) (i-Pr2Ph-DAB) to give thermally unstable 1:1 coordination complexes R2Zn(i-Pr2Ph-DAB), which subsequently undergo a slow regioselective alkyl or aryl group-transfer reaction from the zinc atom to an imine-nitrogen or a carbon atom of the NCCN system of the i-Pr2Ph-DAB ligand. In the case of R=methyl, n-propyl, n-butyl, s-butyl, neopentyl and benzyl, C-alkylation occurs with a subsequent 1,2- hydrogen shift in the amino-imino skeleton affording RZn[(i-Pr2Ph)N-CH2-CR=N(i-Pr2Ph)], whereas for R=t- butyl the C-alkylated product t-BuZn[(i-Pr2Ph)N---CH(t-Bu)---CH=N(i-Pr2Ph)] is stable. Surprisingly, diphenylzinc reacts with i-Pr2Ph-DAB exclusively to give the N-arylated product PhZn[(i-Pr2Ph)N=CHCH=N(Ph)(i-Pr2Ph)].  相似文献   

13.
By reaction of Zn4O(O2CNMe2)6 (1) with [NH2Me2][O2CNMe2] in toluene as medium, the homoleptic zinc compound [Zn(O2CNMe2)2] (2) was obtained, which reverted back to the tetranuclear μ-oxo derivative by controlled hydrolysis. The reaction of ZnO in MeCN with an excess of [NH2Me2][O2CNMe2] in concentrated solution produced high yields of [Zn(O2CNMe2)2] (2) or [NH2Me2][Zn2(O2CNMe2)5] · xMeCN, 3 · xMeCN, x = 1 or 2, depending on the experimental conditions.  相似文献   

14.
The photolysis of [Fe(Et2dtc)3], Et2dtc = diethyldithiocarbamate to yield [Fe(Et2dtc)2Cl] proceeds under 313 nm irradiation through a metal complex excited state, as expected. Under 254 nm irradiation, however, the dominant pathway is through a solvent-initiated reaction in which radicals formed after absorption of light by CHCl3 react thermally with [Fe(Et2dtc)3]. The initial rate varies linearly with the light intensity at 313 nm, but at 254 nm varies with the square root of the intensity.  相似文献   

15.
Single muscle fibers continue to twitch for up to 20 min when immersed in ethylene glycol bis(β-aminoethyl ether)-N,N′-tetraacetic acid (EGTA) solutions containing less than 10−8 M free calcium. Failure of the twitch results from reversible depolarization, which occurs after 15–20 min in EGTA. The results make it clear that external calcium or calcium in the transverse tubules play no essential part in action potential propagation or excitation-contraction coupling.  相似文献   

16.
The modification reaction of sago starch with succinic anhydride (SA) using pyridine (PY) and/or 4-dimethyaminopyridine (DMAP) as catalyst and N,N-dimethylacetamide (DMA)/lithium chloride (LiCl) system as solvent was studied. A series of succinylated starch derivatives were prepared with a degree of substitution (DS) ranging from 0.14 to 1.54. The structure of the resulting polymers determined by means of 13C NMR spectroscopy indicated that substitution preferably occurs at the C2 and C6 hydroxyl groups. The thermal stability of the material was decreased by chemical modification. Effects of reactant molar ratio, reaction time, and the concentrations of DMAP and LiCl on the reaction efficiency are discussed.  相似文献   

17.
18.
Steroid 21-hydroxylase activity has been identified in many tissues, including liver. But it is possible that the enzyme found in the liver is different from adrenal 21-hydroxylase. In the adrenal cortex, steroid 21-hydroxylase activity is increased by corticotropin (ACTH); the effect of ACTH is mediated by cyclic AMP (cAMP), and presumably involves a cAMP-dependent protein kinase (PKA). It is not yet clear, however, how extra-adrenal steroid 21-hydroxylase activity is regulated. In the present study, we examined the effect of N6, 2′-O-dibutyryl adenosine 3′,5′-cyclic monophosphate (dbcAMP), forskolin, N-[2-(methylamino)ethyl]5-isoquinolinesulfonamide (H-8) and 12-O-tetradecanoylphorbol-13-acetate (TPA) on steroid 21-hydroxylase activity in primary cultures of rat hepatocytes to determine the nature of regulation of extra-adrenal steroid 21-hydroxylase activity. Steroid 21-hydroxylase activity in hepatocytes incubated with 10−11M dbcAMP for 24 h was 1.6 times higher than that in control hepatocytes untreated with dbcAMP. On the other hand, steroid 21-hydroxylase activity decreased by 20 and 50% when the cells were incubated with 10−5 and 10−3 M dbcAMP, respectively. The stimulatory effect of 10−11 M dbcAMP was not blocked by 10−5 M H-8 (PKA inhibitor), but the inhibitory effect of 10−5 or 10−3 M cAMP was. TPA did not alter the activity of steroid 21-hydroxylase. These findings indicate that the steroid 21-hydroxylase in rat liver is regulated by mechanisms different from those in the adrenal glands.  相似文献   

19.
Oxygenation of [CuII(fla)(idpa)]ClO4 (fla=flavonolate; IDPA=3,3′-iminobis(N,N-dimethylpropylamine)) in dimethylformamide gives [CuII(idpa)(O-bs)]ClO4 (O-bs=O-benzoylsalicylate) and CO. The oxygenolysis of [CuII(fla)(idpa)]ClO4 in DMF was followed by electronic spectroscopy and the rate law −d[{CuII(fla)(idpa)}ClO4]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2] was obtained. The rate constant, activation enthalpy and entropy at 373 K are kobs=6.13±0.16×10−3 M−1 s−1, ΔH=64±5 kJ mol−1, ΔS=−120±13 J mol−1 K−1, respectively. The reaction fits a Hammett linear free energy relationship and a higher electron density on copper gives faster oxygenation rates. The complex [CuII(fla)(idpa)]ClO4 has also been found to be a selective catalyst for the oxygenation of flavonol to the corresponding O-benzoylsalicylic acid and CO. The kinetics of the oxygenolysis in DMF was followed by electronic spectroscopy and the following rate law was obtained: −d[flaH]/dt=kobs[{CuII(fla)(idpa)}ClO4][O2]. The rate constant, activation enthalpy and entropy at 403 K are kobs=4.22±0.15×10−2 M−1 s−1, ΔH=71±6 kJ mol−1, ΔS=−97±15 J mol−1 K−1, respectively.  相似文献   

20.
The triazenide complex of Pt(II) trans-(o-Tol)Pt(PEt3)2N3Ar2(1) (Ar = p-FC6H4) was synthesized by reaction of (o-Tol)Pt(PEt3)2BF4 with Ar2N3Na. The 1H, 19F and 31P NMR spectra of this complex in toluene-d8 were studied at different temperatures. Two kinds of dynamic processes were observed. The first one is the intramolecular N,N′ migration of the (o-Tol)Pt(PEt3)2 group, detected by 19F NMR. The second process, revealed by 1H, 19P NMR, is the rotation around the partially double N(2)–N(3) bond. Thermodynamic parameters for these processes were calculated from dynamic NMR spectra.  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