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.
Six new dinuclear complexes, derived from cis-[Co(H2O)2(NH3)4]3+, cis-[Co(H2O)2(en)2]3+ and [M(CN)42? (M = Ni, Pd, Pt) were prepared and characterized by means of chemical analysis, electronic and IR measurements. The influence of the pH on the rate of the reaction was studied for the two derivatives of [Pd(CN)4]2?, showing that the best conditions to obtain the dinuclear compounds are at pH near 6, where the predominant species are cis-[Co(OH)(H2O)(amine)2]2+. The [Pt(CN)4]2? derivatives show PtPt interactions both in the solid state and in solution.  相似文献   

2.
Abstract

The 1H NMR relaxation effects produced by paramagnetic Cr(III) complexes on nucleoside 5′-mono- and -triphosphates in D2O solution at Ph′=3 were measured. The paramagnetic probes were [Cr(III)(H2O) 6]3+, [Cr(III)(H2O)3 (HATP)], [Cr(III)(H2O)3(HCTP)] and [Cr(III) (H2O)3(UTP)?, while the matrix nucleotides (0.1 M) were H2AMP, HIMP?, and H2ATP2-. For the aromatic base protons, the ratios of the transverse to longitudinal paramagnetic relaxation rates (R2p/R1p) for the [Cr(III)(H2O)6]3+/H2ATP2-, [Cr(III)(H2O)3(HATP)]/H2ATP2-, [Cr(III)(H2O)3(HCTP)]/H2ATP2 and [Cr(III)(H2O)3(UTP)]?/H2ATP2 systems were below 2.33 so the dipolar term predominates. For a given nucleotide, R1p for the purine H(8) signal was larger than for the H(2) signal with the [Cr(III)(H2O)6]3+ probe, while R1p for the H(2) signal was larger with all the other Cr(III) probes. Molecular mechanics computations on the [Cr(III)(H2O)4(HPP)(α,β)], [Cr(III)(NH3)4(HPP)(α,β)], [Co(III)(NH3)3(H2PPP)(α,βγ)] and [Co(III)(NH3)4(HPP)(α,β)] complexes gave calculated energy-minimized geometries in good agreement with those reported in crystal structures. The molecular mechanics force constants found were then used to calculate the geometry of the inner sphere [Cr(III)(H2O)6]3+ and [Cr(III)(H2O)3(HATP)(α,βγ)] complexes as well as the structures of the outer sphere [Cr(III) (H2O)6]3+-(H2AMP) and [Cr(III)(H2O)6]-(HIMP)? species. The gas-phase structure of the [Cr(III)(H2O)3(HATP)(α,βγ)] complex shows the existence of a hydrogen bond interaction between a water ligand and the adenine N(7) (O…N = 2.82 Å). The structure is also stabilized by intramolecular hydrogen bonds involving the -O(2′)H group and the adenine N(3) (O…N = 2.80 Å) as well as phosphate oxygen atoms and a water molecule (O…O = 2.47 Å). The metal center has an almost regular octahedral coordination geometry.

The structures of the two outer-sphere species reveal that the phosphate group interacts strongly with the hexa-aquochromium probe. In both complexes, the nucleotides have a similar “anti” conformation around the N(9)-C(l′) glycosidic bond. However, a very important difference characterizes the two structures. For the (HIMP)? complex, strong hydrogen bond interactions exist between one and two water ligands and the inosine N(7) and O(6) atoms, respectively (O…O = 2.63 Å O…N = 2.72, 2.70 Å). For the H2AMP complex, the [Cr(III) (H2O)c]3+ cation does not interact with N(7) since it is far from the purine system. Hydrogen bonds occur between water ligands and phosphate oxygens. The Cr-H(8) and Cr-H(2) distances revealed by the energy-minimized geometries for the two outer sphere species were used to calculate the R1p values for the H(8) and H(2) signals for comparison with the observed R1p values: 0.92(c), 1.04(ob) (H(8)) and 0.06(c), 0.35(ob) (H(2)) for H2AMP; and 3.76(c), 4.53(ob) (H(8)) and 0.16(c), 0.77(ob) s?1 (H(2)) for HIMP?. These results suggest that the dynamic relaxation effects can be only partially understood with molecular mechanics computations, although the success of the geometry calculations suggests that future efforts in the development of computational methods are justified.  相似文献   

3.
Monodentate Co(NH3)5PPi was determined not to be a substrate for yeast inorganic pyrophosphatase while P1,P2-bidentate Co(NH3)4PPi was turned over by the enzyme at a rate of 7.5 min?1. A kinetic analysis of the substrate activities of the P1,P2-bidentate complexes, Co(en)2PPi, Cr(NH3)4PPi, Cr(H2O)(NH3)3PPi, Cr(H2O)2(NH3)2PPi, and Cr(H2O)4PPi was carried out in order to access the potential role of the metal-water ligands in productive binding. While substitution of the H2O ligands with NH3 ligands had a minimal affect on the Km for Mg2+, the binding affinity of the complexes decreased with an increasing NH3H2O ligand ratio as did the turnover number of the corresponding central complexes. The Co(en)2PPi complex was hydrolyzed at a rate approximately 0.6% of that for the Co(NH3)4PPi complex. The substrate activities of β,γ-bidentate Co(NH3)4PPPi and α,β,γ-tridentate Co(NH3)3PPP with pyrophosphatase were also tested. While both complexes were shown to bind tightly to the Mg2+-activated enzyme neither was hydrolyzed. On the other hand, in the presence of the Zn2+-activated enzyme the tridentate complex was turned over at a rate of 0.17 min?1 while the bidentate complex remained inert to hydrolysis.  相似文献   

4.
Products of the reduction of [CoNO2(NH3)5]2+ by Cr2+ were separated and identified under the conditions of [Cr2+]0/[Co(IlI)]0⩽3 and 0.02 M ⩽[H+] ⩽ 0.75 M. The product distribution was dependent on both [Cr2+]o and [H+]. The following mechanism is proposed: [CoNO2(NH3)5]2+ + Cr2+→Co2+ + [CrONO(H2O)5]2+ (i) [CrONO(H2O)5]2+ + H+→[Cr(H2O)6]3+ + HNO2 (ii) [CrONO(H2O)5]2+ + Cr2+→Cr(IV) + [CrNO(H2O)5]2+ (iii) Cr(IV) + Cr2+→[(H2O)4Cr(OH)2Cr(H2O)4]4+ (iv) HNO2 + 2Cr2+→[Cr(H2O)6]3+ + [CrNO(H2O)5]2+ (v)  相似文献   

5.
The reaction of [(H2O)(NH3)5RuII]2+ with calf thymus and salmon sperm DNA has been studied over a wide fange of transition metal ion concentrations. Kinetic studies revealcd a biphasic reaction with an initial fairly rapid coordination of the metal ion being followed by slower reactions. Binding studies were done under pseudo-equilibrium conditions following completion of the initial rapid reaction. Spectra and HPLC of acid-hydrolyzed samples of [(NH3)5RuII]n-DNA prepared by incubation of [(H2O)(NH3)5RuII]2+ with DNA (where [PDNA] = 1.5 mM and reactant [RuII]/[PDNA] ratios were in the fange 0.1 to 0.3) followed by air oxidation showed the predominant binding site on helical DNA to be in the major groove at the N-7 of guanine. The equilibrium constant for [(H2O)(NH3)5RuII]2+ binding to the G7 site in helical CT DNA is 5.1 x 103. Differential pulse voltammetry exhibited a single peak at 48 mV, which is attributed to the reduction of Rum on the G7 sites.At [Run]/[PDNA] <0.5, Tm values for the DNA decreased linearly with increasing ruthenium concentration and an increase in the intensity of the 565 nm dG→ Ru(III) charge transfer band was noted upon melting. The UV and CD spectra of these samples indicated no extensive destacking or alteration in geometry (B family) compared to unsubstituted DNA. At [Run]/[PDNA]〉 0.5 or when single-stranded DNA was used, increased absorbance at 530 nm and 480 nm suggested additional binding to the exocyclic amine sites of adenine and cytosine residues. HPLC and individual spectrophotometric identification of the products derived from hydrolysis of these spec~es yielded both [(Gua)(NH3)5RuIII] and [(Ade)(NH3)5RuIII]. Earlier studies have established the cytidine and adenosine binding sites of [(NH3)5RuIII] to be at their exocyclic amines (C4 and A6). Coordination to these positions indicates disruption of the double helix since these amines are involved in hydrogen bonding on the interior of B-DNA.Agrose gel electrophoresis of superhelical pBR322 plasmid DNA after exposure to various complexes of [(nh3)5Ruiii] in the presence of a reductant and air generally revealed moderately efficient cleavage of the DNA, presumably due to the generation of hydroxyl radical via Fenton's chemistry. However, similar studies involving [(NH3)5RuIII] directly coordinated to the DNA showed no strand cutting above background. Polyacrylamide gel electrophoresis of a 381 bp, 3′-32P-labeled fragment of pBR322 plasmid DNA containing low levels of bound [(NH3)5 RuIII] further indicated negligible DNA cutting by the coordinated metal ion.  相似文献   

6.
The cationic one-dimensional (1D) coordination polymer chain 1{[Co(μ-bpdo)(H2O)4]2+} and the metal-complex anion trans-[Co(SO4)2(bpdo)2(H2O)2]2−, both based on the 4,4′-bipyridine-N,N′-dioxide (bpdo) ligand, form a complementary supramolecular pair 1{[Co(μ-bpdo)(H2O)4]2+}nn[Co(SO4)2(bpdo)2(H2O)2]2− (1) with respect to charge balance and hydrogen bonding. With a length of >22.14 ? along the bpdo-Co-bpdo axis the metal-complex trans-[Co(SO4)2(bpdo)2(H2O)2]2− is one of the longest and anisotropic counter anions (aspect ratio 22.14:8.11:4.17) observed so far in coordination polymers. Hydrogen-bonding of the anion links the cationic metal-organic 1D polymer into a 2-fold interpenetrated three-dimensional (3D) fsc (or sqc11) 4,6-c 2-nodal net of stoichiometry (4-c)(6-c) with square-planar, 4-connected (Co in anion) and octahedral, 6-connected (Co in cation) nodes in a 1:1 ratio. The 4-c point symbol is (44.62), the 6-c one (44.610.8) yielding a point symbol for the fsc net of (44.62)(44.610.8). The synthesis of 1 requires the presence of a Schiff base. Synthesis under the same conditions in the absence of the Schiff base yields the molecular complex and cocrystal [Co(bpdo)(H2O)5]SO4·1/2bpdo (2) which is related (as pseudo-polymorph) to the known solvate [Co(bpdo)(H2O)5]SO4·2H2O (3) (CSD Refcodes RAXMUZ and RAXMUZ01).  相似文献   

7.
Reactions of CH3[Co] with (CH3)nM(4?n)+ (n = 2, 3; M = Sn, Pb) at concentrations high enough to detect (CH3)4M in the head space (yields 7.08×10?5?2.06×10?5%), indicate that dismutation is the major route of production. Similarly, kinetic reactions at lower concentrations show that no demethylation of CH3[Co] by (CH3)3M+ (M = Sn, Pb) occurs after 60 days. From the methylation of SnCl2 by CH3[Co] at pD 1.0 and under aerobic conditions, the following hydrolysis species were observed in the 400 MHz 1H NMR spectrum: CH3- Sn(OH)Cl2·2H2O (63.6%), [CH3Sn(OH)(H2O)4]2+ (17.6%) and CH3Sn(OH)2Cl·nH2O (18.8%). No methylation products were observed from similar reactions with Pb(II) salts.  相似文献   

8.
The uranium(IV) complexes [U(EDTA)(H2O)2], [U(HOEDTA)]+, and [U(DTPA)]? are well-formed in the pH fange 2–3 ([DTPA]5- = diethylenetriaminepentaacetate; [HOEDTA]3-  N-(2-hydroxyethyl)ethylenediaminetriacetate). Of these, only [U(DTPA)]- is extracted from an aqueous phase at pH 2 by the perchlorate salt of the primary amine, Primene JM-T. As the aqueous phase pH was raised, extraction occurred in all three cases and hydrolysed species may be extracted from EDTA and HOEDTA solutions but [U(DTPA)]? resists hydrolysis. The addition of sulphate had a marked effect on the extraction of U(IV) from EDTA and HOEDTA through the formation of [U(EDTA)(SO4)(H2O)]2- and [U(HOEDTA)(SO4)(H2O)n]?. The equilibrium constant, log β1, for: [(U(EDTA)(H2O)2] 2 [SO4]2? ? [U(EDTA)(SO4)(H2O)]2- 2 H2O was found to be 2.43 ± 0.04 (I = 1 mol dm?3, NaClO4; pH 2.0; 20 °C) from spectrophotometric data.With tri-n-octylphosphine oxide (TOPO) electronic spectroscopy showed that the same U(IV) complex was extracted at pH 2 for Cs2UCl6, U(IV)/ HOEDTA, and U(IV)/DTPA and the aminepoly- carboxylates were aqueous phase masking agents but with [U(EDTA)(H2O)2] oxidation gave a uranyl(VI) organic phase species.Uranium(IV) is strongly extracted from aqueous solutions of HOEDTA at pH 2 or 3 by bis(2-ethyl- hexyl)phosphoric acid (HBEHP) but less so from EDTA and DTPA. Since U(IV) is completely extracted from Cs2UCl6 it could be that the amine- polycarboxylates were aqueous phase masking agents although spectral evidence did not support this.  相似文献   

9.
The salts - yellow [Cr(NH3)6][Ag(CN)2]3 · 2H2O, red [Co(NH3)6][Ag(CN)2]3 · 2H2O, red [Co(NH3)6][Au(CN)2]3 · 2H2O, pale yellow [Ru(NH3)6][Ag(CN)2]3 · 2H2O, yellow K[Cr(NH3)6]2[Au(CN)2]7 · 4H2O, and colorless [(μ2-NH2)2Pt2(NH3)10][Au(CN)2]6 · 5.5{OS(CH3)2} · 0.5H2O - have been prepared by evaporation of aqueous solutions of potassium dicyanoargenate or potassium dicyanoaurate and salts of the appropriate cations. Hydrogen bonding between the cations and the cyano groups of the anions facilitates the formation of structures with strong metallophilic interactions between the anions. Thus, the [Au(CN)2] or [Ag(CN)2] ions self-associate into linear trimers in the isostructural set of crystals, [Cr(NH3)6][Ag(CN)2]3 · 2H2O (Ag?Ag distance; 3.1610(4) Å), [Co(NH3)6][Ag(CN)2]3 · 2H2O (Ag?Ag distance; 3.1557(2) Å), [Co(NH3)6][Au(CN)2]3 · 2H2O (Au?Au distance; 3.0939(4) Å), and [Ru(NH3)6][Ag(CN)2]3 · 2H2O (Ag?Ag distance; 3.1584(5) Å). Crystalline [(μ2-NH2)2Pt2(NH3)10][Au(CN)2]6 · 5.5{OS(CH3)2} · 0.5H2O also contains nearly linear trimers of the dicyanoaurate ion. Yellow crystals of K[Cr(NH3)6]2[Au(CN)2]7 · 4H2O contain a centrosymmetric, bent chain of seven dicyanoaurate ions with Au?Au separations of 3.1806(3), 3.2584(4), and 3.1294(4) Å.  相似文献   

10.
The mixed diamine complexes trans-[Co(tmen)(diamine)Cl2]+ have been synthesised (tmen = NH2C(Me)2C(Me)2NH2; diamine = en = NH2(CH2)2NH2, and ibn = NH2C(Me)2CH2NH2). Replacement of one en ligand in trans-[Co(en)2Cl2]+ by one tmen ligand engenders an enormous rate enhancement (2000-fold) for acid hydrolysis. Solvolysis rates have been measured in Me2SO and DMF for these complexes and also trans-[Co(tmen)2Cl2]+ which is more reactive again (104-fold). The measured reactivities in DMF at 2 °C establish that the kinetic effect of replacing each en by tmen is incremental, and the extreme base catalysed racemisation rate for (+)-[Co(tmen)3]3+ can now be explained on this basis.  相似文献   

11.
 The synthesis, spectroscopic, and electrochemical properties of trans-[L(Pyr)(NH3)4RuII/III] (Pyr=py, 3-phpy, 4-phpy, 3-bnpy, or 4-bnpy; L=H2O, Guo, dGuo, 1MeGuo, Gua, Ino, or G7-DNA) are reported. As expected, the Pyr ligand slows DNA binding by trans-[(H2O)(Pyr)(NH3)4RuII]2+ relative to [(H2O)(NH3)5RuII]2+ and favors reduction of RuIII by about 150 mV. The pyridine ligand also promotes the disproportionation of RuIII to afford the corresponding complexes of RuII and, presumably, RuIV. For L=Ino, disproportionation follows the rate law: d[RuII]/dt=k 0[RuIII]+k 1[OH][RuIII], k 0=(2.7±0.7)×10–4 s–1 and k 1=70±1 M–1 s–1. Received: 11 May 1998 / Accepted: 3 March 1999  相似文献   

12.
Reaction between [(C5H5)Co{P(O)(OEt)2}3]2UCl2 and neopentyl lithium affords the novel complex, [{η4-C5H5(CH2C(CH3)3)}Co{P(O)(OEt)2}3]2U, in which the uranium metal center has been dehalogenated and the neopentyl nucleophiles have attacked the cyclopentadienyl groups on the Kläui ([(C5H5)Co{P(O)(OEt)2}3]) ligands. The uranium atom in the title compound possesses octahedral geometry defined by the oxygen atoms from two sets of tripodal oxygen ligands, while the cyclopentadienyl ligands are bound η4 to the cobalt atoms. The formation of this complex suggests that the Kläui ligand may not be a suitable ligand framework for supporting organometallic complexes of oxophilic early actinides.  相似文献   

13.
The interaction of guanine, guanosine or 5-GMP (guanosine 5-monophosphate) with [Pd(en)(H2O)2](NO3)2 and [Pd(dapol)(H2O)2](NO3)2, where en is ethylenediamine and dapol is 2-hydroxy-1,3-propanediamine, were studied by UV-Vis, pH titration and 1H NMR. The pH titration data show that both N1 and N7 can coordinate to [Pd(en)(H2O)2]2+ or [Pd(dapol)(H2O)2]2+. The pKa of N1-H decreased to 3.7 upon coordination in guanosine and 5-GMP complexes, which is significantly lower than that of ∼9.3 in the free ligand. In strongly acidic solution where N1-H is still protonated, only N7 coordinates to the metal ion, but as the pH increases to pH ∼3, 1H NMR shows that both N7-only and N1-only coordinated species exist. At pH 4-5, both N1-only and N1,N7-bridged coordination to Pd(II) complexes are found for guanosine and 5-GMP. The latter form cyclic tetrameric complexes, [Pd(diamine)(μ-N1,N7-Guo]44+ and [Pd(diamine)(μ-N1,N7-5-GMP)]4Hx(4−x)−, (x=2,1, or 0) with either [Pd(en)(H2O)2](NO3)2 or [Pd(dapol)(H2O)2](NO3)2. The pH titration data and 1H NMR data agree well with the exception that the species distribution diagrams show the initial formation of the N1-only and N1,N7-bridged complexes to occur at somewhat higher pH than do the NMR data. This is due to a concentration difference in the two sets of data.  相似文献   

14.
Abstract

The interaction of adenosine-5′-monophosphate (5′-AMP), guanosine-5′-monophosphate (5′-GMP) and 2′-deoxyguanosine-5′-monophosphate (5′-dGMP) with the [Co(NH3)6]3+, [CO(NH3)5C1]2+ and [CO(NH3)4C12]+ cations has been investigated in aqueous solution with metal/nucleotide ratios (r) of 1/2, 1 and 2 at neutral pH. The solid complexes have been isolated and characterized by FT-IR and 1H-NMR spectroscopy.

The complexes are polymeric in nature both in the crystalline solid and aqueous solution. The binding of the cobalt-hexammine cation is indirectly (via NH3) through the N-7 and the PO3 2- groups of the AMP and via O-6, N-7 and the PO3 2- of the GMP and dGMP anions (outer-sphere). The cobalt-pentammine and cobalt-tetrammine bindings are through the phosphate groups (inner-sphere) and the N-7 site (outer-sphere) of these nucleotide anions. The ribose moiety shows C2′-endo/anti conformation, in the free AMP and GMP anions as well as in the cobalt-ammine - AMP complexes, whereas a mixture of the C2′-endo/anti and C3′-endo/anti sugar puckers were observed for the Co(NH3)6-GMP, Co(NH3)5-GMP and a C3′-endo/anti conformer for the Co(NH3)4-GMP complexes. The deoxyribose showed an O4′-endo/anti conformation for the free dGMP anion and a C3′-endo/anti for the Co(NH3)6-dGMP, Co(NH3)5-dGMP and Co(NH3)4-dGMP complexes.  相似文献   

15.
《Inorganica chimica acta》1988,150(1):81-100
The (NH3)5CoOC(NH2)23+ ion is consumed in water according to the rate law k(obs.) = k1 + k2[OH], where k1 = 4.0 × 10−5 s−1 and k2 = 14.2 M−1 s−1 (0–0.1 M [OH];μ = 1.1 M, NaClO4, 25 °C). A hitherto unrecognized intramolecular O- to N- linkage isomerization reaction has been detected. In strongly acid solution only aquation to (NH3)5CoOH23+ is observed, but in 0.1–1.0 M [OH], 7% of the directly formed products is the urea-N complex (NH3)5CoNHCONH22+ which has been isolated. In the neutral pH region a much greater proportion (25%) of the products is the urea-N species. These results are interpreted in terms of an urea-O to urea-N linkage isomerization reaction competing with hydrolysis for both spontaneous (k1) and base-catalyzed (k2) pathways; the rearrangement is not observed in strongly acidic solution (pH ⩽ 1) because the protonated N-bonded isomer (pKa ≈ 3) is unstable with respect to the O-bonded form. The appearance of the isomerization pathway as the pH is raised in the 0–6 region is commensurate with a rate increase which cannot be attributed to a contribution from the base catalysis term k2[OH]. It is argued that this observation establishes, for the spontaneous pathway, that hydrolysis and linkage isomerization are separate reaction pathways — there is no common intermediate. The product distribution and rate data lead to the complete rate law, k(obs.) = k1 + k2[OH] = (ks + kON) + (kOH + kON) [OH] for the reactions of the O-bonded isomers, where ks, kOH are the specific rates for hydrolysis, and kON, kON are the specific rates for O- to N-linkage isomerization, by spontaneous and base-catalyzed pathways respectively; kON = 1.3 × 10−5 s−1 and kON = 1.1 M−1 s−1 (μ = 1.0 M, NaClO4, 25 °C). The O- to N- linkage isomerization has been observed also for complexes of N-methylurea, N,N-dimethylurea and N-phenylurea, but not for the N,N′-dimethylurea species. There is an approximately statistical relationship among the data for −NH2 capture (versus H2O), while −NHR and −NR2 do not compete with water as nucleophiles for Co(III) in either the spontaneous or base-catalyzed hydrolysis processes. For each urea-O complex, O- to N-isomerization is a more significant parallel reaction in the spontaneous as opposed to the base-catalyzed pathway. This is interpreted as being indicative of more associative character in the spontaneous route to products, a conclusion supported by other evidence. Some activation parameter data have been recorded and the effect of the N-substitution on the rates of solvolysis (H2O, Me2SO) is discussed. The urea-N complexes have been isolated as their deprotonated forms, [(NH3)5CoNHCONRR′](ClO4)2·xH2O (R,R′ = H, CH3). They are kinetically inert in neutral to basic solution but in acid they protonate (H2O, pKa 2–3; μ = 1.0 M, 25 °C) and then isomerize rapidly back to their O-bonded forms. Some solvolysis accompanies this N- to O-rearrangement in H2O and Me2SO. Specific rates and activation parameters are reported. The kinetic data follow a rate law of the form kNO(obs.) = (k + kNO)[H+]/(Ka + [H+]) and the active species in the reaction is the protonated form; k, kNO are the specific rates for hydrolysis and isomerization, respectively. Proton NMR data establish that the site of protonation (in Me2SO) is the cobalt-bound nitrogen atom. For the unsubstituted urea species (NH3)5CoNH2CONH23+, diastereotopic exo-NH2 protons arising from restricted rotation about the CN bond are observed. The relevance to the mechanism of the linkage isomerization process is considered. 13C and 1H NMR and electronic absorption spectral data are presented, and distinctions between linkage isomers and the solution structures (electronic and conformational) are discussed. The urea-N/urea-O complex equilibrium is governed by the relation KNO(obs.) = KNO[H+]/[H+](Ka), where KNO is the equilibrium constant = [(NH35Co(urea-O)3+]/[(NH3)5Co(urea-N)3+]. Values for KNO(=kNO/kON = 260 and pKa ≈ 3 for the NH2CONH2 system are consistent with the stability of the N-isomer in feebly acidic to basic solution (e.g. pH 6, KNO(obs.) = 2.6 × 10−2) and instability in acid solution (e.g. pH 1, KNO(obs.) = 240). The equilibrium data for this and other urea complexes of (NH3)5Co(III) are contrasted with the result for the analogous Rh(III)NH2CONH2 system KNO ≈ 1).  相似文献   

16.

The hydrolysis of cyclic adenosine 3′,5′-monophosphate and 2′-deoxythymidylyl(3′-5′)2′-deoxythymidine by Ce(NH4)2(NO3)6 was kinetically studied. The rate of hydrolysis was fairly proportional to the concentration of [Ce IV 2 (OH)4]4+, showing that this is the catalytically active species. According to quantum-chemical calculation, the two Ce(IV) ions in this [CeIV 2(OH)4]4+ cluster are bridged by two OH residues. Upon the complex formation with H2 PO4 ? (a model compound for the phosphodiesters), these two Ce(IV) ions bind the two oxygen atoms of the substrate and enhance the electrophilicity of the phosphorus atom. The catalytic mechanism of Ce(IV)-induced hydrolysis of phosphodiesters has been proposed on the basis these results.  相似文献   

17.
The interactions between N-tosylamino acids and cobalt(II), nickel(II) and zinc(II) ions in aqueous solution and in the solid state have been investigated. From concentrated aqueous solutions, compounds of general formula [M(II)(N-tosylaminoacidato)2(H2O)4](M = Co(II), Ni(II) and N-tosylaminoacidato = N-tosylglycinate (Tsgly?), N-tosyl-α- and -β-alaninate (Ts-α- and Ts-β-ala?); M = Zn(II) and N-tosylaminoacidate = Tsgly?, Ts-β-ala?) and [Zn(II)(N- tosylaminoacidato)2(H2O)2] were isolated and characterized by means of thermogravimetric, electronic and infrared spectra. For two of them: [Co(Tsgly)2(H2O)4](I) and [Zn(Ts-β-ala)2(H2O)4](II) the crystal and molecular structures were also determined. Both compounds crystallize in the monoclinic space group P21/c, with two formula units in a cell of dimensions: a = 13.007(6), b = 5.036(2), c = 18.925(7) Å, β = 102.33(3)° for (I) and a = 14.173(6), b = 5.469(2), c = 17.701(7) Å, β = 106.63(3)° for (II). The structures were solved by the heavy-atom method and refined by least-squares calculations to R = 0.031 and 0.064 for (I) and (II) respectively. The cobalt and zinc atoms lie in the centers of symmetry, each bonded to two amino- acid molecules through a carboxylic oxygen atom and four water molecules in a slightly tetragonally distorted octahedral geometry. The second carboxylic oxygen atom is not involved in metal coordination. Electronic and X ray-powder spectra suggest that the tetrahydrate complexes of Co2+, Ni2+ and Zn2+ ions of the same amino acids are isomorphous and isostructural. No coordinative interactions between ligand and metal ions were found in aqueous solution on varying the pH values before hydroxide precipitation.  相似文献   

18.
Red or orange crystals of [Co(NH3)6]2Cl2[Fe(CN)6] · 4H2O (1), [Co(en)3]2Cl2[Fe(CN)6] · 2H2O (2) and [Co(en)3]4[Fe(CN)6]3 · 21.6H2O (3) were isolated from the aqueous systems Co3+-LN-[Fe(CN)6]4− (LN = NH3, en = 1,2-diaminoethane). In all isolated samples the combination of Mössbauer (δ values were from the range −0.07 to −0.08 mm/s) and IR spectra (ν(CN) stretching vibrations in the range 2015-2047 cm−1) confirms the presence of low spin Fe(II) in [Fe(CN)6]4− anions. X-ray structure analyses corroborate the ionic character of all studied compounds. These contain diamagnetic [Co(NH3)6]3+ (1) or [Co(en)3]3+ (2 and 3) complex cations and diamagnetic [Fe(CN)6]4− complex anions. In compounds 1 and 2 chloride anions are present, too. All three compounds contain water of crystallization, in compound 3 as many as 21.6 molecules per formula unit.  相似文献   

19.
Rate parameters have been obtained for the oxidation of cuprous stellacyanin by cobalt(III) ions of the form cis(N)-[CoN2O4]?, including cis(N)-[Co(NTA)(gly)]?, cis(N)-[Co(IDA)2]?, [Co(en)(ox)2]?(μ 0.5 M(phosphate), pH 7.0), and Co(EDTA)?(μ 0.1 M(NaCl), pH 7.2, 0.001 M phosphate). An excellent isokinetic correlation between the activation parameters ΔH and ΔS exists for the reactions of aminopolycarboxylatocobalt(III) ions with reduced stellacyanin (β = 300 ± 12 K; correlation coefficient = 0.995). It is concluded that enthalpy-entropy compensation in these reactions may be understood in terms of differing orientations preferred by the various oxidants in forming precursor complexes with the reduced blue protein. While ΔH and ΔS values for electron transfer from stellacyanin to cis(N)-[CoN2O4]? ions vary over ranges of 10.7 kcal/mol and 34 cal/mol-deg, respectively, room temperature rate constants are relatively constant (3.6–34.5 M?1 sec?1), as expected from Marcus theory for outer sphere electron transfer.  相似文献   

20.
Transition metal complexes [Co(cyclen)(NH3)2](ClO4)3⋅H2O (cyclen = 1,4,7,10-tetraazacyclododecane) (2), [Co(NH3)5(OH2)](CF3SO3)3 (3) [Ni(NH3)6]Br2 (4) and [Ru(NH3)6]Cl3 (5) were tested against Sindbis infected baby hamster kidney (BHK) cells and show differential effects from the previously reported anti-viral complex [Co(NH3)6]Cl3 (1). The macrocyclic complex 2 and labile aqua complex 3 show either no or little effect on the survival on Sindbis virus-infected cells as compared to that for 1, which show a monotonic increase in % BHK cell survival. Nickel and ruthenium ammine complexes 4 and 5 had a moderate influence of cell survival. While the results showed some anti-viral activity for some of the structural variations, it appears that 1, with its potential to be a broad-spectrum anti-viral compound, occupies a unique position in its ability to both significantly enhance cell survival and to decrease viral expression of infected cells. We also show that 1 also shows anti-viral activity against Adenovirus lending support to the broad-spectrum potential of this complex.  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