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 stability constants of the Ni2+ and Co2+ complexes with 1,5-diazacyclooctane-N,N′-diacetic acid (H2DACODA) have been determined potentiometrically in 0.5M KNO3 at 25°. Only M(DACODA) and M(DACODA)OH? were observed. In addition the formation and dissociation kinetics of the pentacoordinate complexes M(DACODA) has been studied in aqueous solution using a stopped-flow technique. Formation follows the rate law vf = kf [M2+] [HDACODA?]/[H+], which can be interpreted as a bimolecular process either between M2+ and DACODA2? (k) or between MOH+ and HDACODA? (k). The second order rate constants k are much higher than those expected from water exchange and can only be explained by a strong internal conjugate base effect. In the limiting case, however, this is equivalent to the second possible explanation, which assumes MOH+ and HDACODA? as reactive species. The dissociation rate is given by vd = (kML + k [H+]) · [M(DACODA)].  相似文献   

2.
1-(2′-Deoxy-2′-fluororibofuranosyl)pyrimidines were synthesized and incorporated into an RNA oligonucleotide to give 5′-r[CfGCf(UfUfCfG)GCfG]-3′ (Cf: short form of C = 2′-deoxy-2′-fluorocytidine; Uf: short form of U = 2′-deoxy-2′-fluorouridine). The oligomer was investigated by means of UV, CD, and NMR spectroscopy to address the question of how F-labels can substitute 13C-labels in the ribose ring. Through-space (NOE) and through-bond (scalar couplings) experiments were performed that make use of the ameliorated chemical-shift dispersion induced by 19F as an alternative heteronucleus. A comparison of the structures of fluorinated vs. unmodified oligomer is given. It turns out that the fluorinated oligonucleotide exists in a 14:3 equilibrium between a hairpin and a duplex conformation, in contrast to the unmodified oligonucleotide which predominantly adopts the hairpin conformation. Furthermore, the fluorinated hairpin structure adopts two distinct conformations that differ in the sugar conformation of the U and C nucleoside units, as detected by the 19F-NMR chemical shifts. The role of the 2′-OH group as stabilizing element in RNA secondary structure is discussed.  相似文献   

3.
The ternary Cu2+?2,2′-bipyridyl-adenosine-5′-monophosphate-N(1)-oxide complex was investigated and compared with the binary Cu2+-adenosine-5′-monophosphate-N(1)-oxide complex (I) (cf. [2]). In both complexes Cu2+ is bound to the o-amino-N-oxide group of adenosine-5′-monophosphate-N(1)-oxide (HL). The stabilities of the complexes monoprotonated at the phosphate group are of the same order: log K = 11,20, and log K = 11,19. The acidity constants for the deprolonation of the phosphate group in these complexes are slightly different (pK = 5,55, and pK = 5,88), but as expected both values are lower than the corresponding value pK = 6,12 of the ligand.  相似文献   

4.
Protonation and Cu(II) complexation equilibria of L -phenyhilaninamide, N2-methyl-L-phenylalaninamide, N2, N2-dimethyl-L-phenylalaninamide, L -valinamide, and L -prolinamide have been studied by potentiometry in aqueous solution. The formation constants of the species observed, CuL2+, CuL, CuLH, CuL2H and CuL2H?2, are discussed in relation to the structures of the ligands. Possible structures of bisamidato complexes are proposed on the ground of VIS and CD spectra. Since Cu(II) complexes of the present ligands (pH range 6–8) perform chiral resolution of dansyl- and unmodified amino acids in HPLC (reversed phase), it is relevant for the investigation of the resolution mechanism to know which are the species potentially involved in the recognition process.  相似文献   

5.
Aqueous sols of TiO2 (anatase, particle radius 25 Å) were excited with (347.1 nm)-laser light and the reaction of valence-band holes with halide ions (X = I?, Br?, Cl?) was investigated. Hole transfer takes place within the duration of the (10 ns)-laser pulse and results in the formation of anion radicals according to the sequence: The quantum yield of X increases in the order Cl < Br < I, attaining 0.8 for I at pH 1. It is affected by pH, halide concentration and the presence of a protective agent for the sol. RuO2 deposited onto TiO2 enhances markedly Cl and Br -formation, but has no effect on the yield of I. Laser-photolysis investigation of halide oxidation were also carried out with colloidal Fe2O3 (particle radius 600 Å). For I2?formation, the quantum yield exceeds 0.9 indicating almost quantitative hole scavenging by iodide.  相似文献   

6.
The temperature dependence of the emission lifetime of the series of complexes Ru(bpy)n(4,4′-dpb) (bpy = 2,2′bipyridine, 4,4′-dpb = 4,4′-diphenyl-2,2′-bipyridine) has been studied in propionitrile/butyronitrile (4:5 v/v) solutions in the range 90–293 K. The obtained photophysical parameters show that the energy separation between the metal-to-ligand charge tranfer (3MLCT) emitting level and the photoreactive metal-centered (3MC) level changes across the series (ΔE = 3960, 4100, 4300, and 4700 cm?1 for Ru(bpy)), Ru(bpy)2(4,4′-dpb)2+, Ru(bpy)(4,4′-dpb), and Ru(4,4′-dpb), respectively, where ΔE is the energy separation between the minimum of the 3MLCT potential curve and 3MLCT – 3MC crossing point. Comparison between spectral and electrochemical data indicated that the changes in ΔE are due to stabilization of the MLCT levels in complexes containing 4,4′-dpb with respect to Ru(bpy)2+3. The photochemical data for the same complexes (as I? salts) have been obtained in CH2Cl2 in the presence of 0.01M Cl? upon irradiation at 462 nm. The complexes containing 4,4′-dpb are more photostable than Ru(bpy). Comparison between the data for thermal population of the 3MC photoreactive state and those for photochemistry indicated that the overall photochemical process is governed by (i) a thermal redistribution between the emitting and photoreactive excited states, and (ii) mechanistic factors, likely related to the size of the detaching ligand.  相似文献   

7.
Results are reported for high-energy beam experiments which establish the formation of endohedral carbon cluster-noble gas compounds by bimolecular reactions of C (x = 60, 70; n = 1, 2, 3) with He and C with Ne. The ions were accelerated up to 8 ke V in a four-sector mass spectrometer and allowed to collide with the noble gas in a collision chamber at room temperatur. Product ions were monitored with a B/E = constant linked scan. Within the sensivity of the experiments, no carbon cluster-gas compounds were observed in the reactions of C with H2, D2, O2, Ar, and SF6, or of C with O2. The observed fall in the cross-section for carbon cluster-noble gas compounds with increasing size of the noble gas, the observation of unimolecular loss of C2 from mass-selected CxHe+ ions, and the elimination of carbon fragments instead of He observed in the formation of the collision-induced CxHen+ product ions are taken as evidence for endohedral compound formation. Results of ab initio molecular-orbital calculations for the perpendicular penetration of the plane of ionized benzene with He, Ne, and Ar indicate that sufficient kinetic energy should be available in the collisions with C to penetrate the C cage at the collision energies of the experiments.  相似文献   

8.
63Cu-NMR.-Spectra of Cu(CH3CN)4X (X = ClO, BF, PF) and Cu(C5H5N)4X (X = ClO, BF) in solution are reported at different temperatures and concentrations. The influence of temperature on the linewidth and chemical shift indicates an equilibrium of Cu(CH3CN) and Cu(C5H5N) with another complex of lower symmetry. The preferential solvation of Cu (I) by pyridin in a mixture acetonitrile/pyridine is clearly shown.  相似文献   

9.
The title cation ( = Ni2L) is formed in a variety of reactions (Schemes 1 and 2) in systems containing Ni2+ and (2-thiolatoethyl)-diphenylphosphine (= L?) in the absence of coordinating anions at Ni2+/L? ratios > 0.5 in apolar or moderately polar media. Solid [Ni2L3]CIO4 and [Ni2L3]BPh4 have been isolated. Job's plots confirm the Ni2L- stoichiometry in solution. 31P-NMR data are consistent with ≥ 97% Ni2L (vs. ? 3% of hypothetical Ni3L) at equilibrium and support the suggested configuration (Fig. 2). The equilibrium between NiL2 + NiL2Br2 and Ni2L + Br? varies with the solvent composition in CH23Cl2/EtOH mixtures. The rate of formation of Ni2L2Br2 from Ni2L and bromide (in high excess) in CH2Cl2 is first-order in [Ni2L]tot but depends on the ratio [Bu4NBr]tot/[Ni2L3 · ClO4]tot, even at a high excess of bromide. This is interpreted by efficient competition in ion-aggregate formation between the small perchlorate concentration introduced as the counterion of Ni2L, and the large excess of bromide.  相似文献   

10.
The title compound and its potassium analog have been prepared from corresponding aqueous solutions of 99TcO at pH ≈? 2 with SO2 as a reducing agent. An X-ray structure determination of the Na-salt showed Tc coordinated to the tetradentate N(CH2COO) ligand (NTA). Two Tc-NTA moieties are joined via two bridging O-atoms into a four-membered Tc2O2 ring. The observed diamagnetism, a strong absorption band at 19 950 cm?1, and a short Tc-Tc distance of 2.363 Å are typical for the Tc2O2-fragment with its strong metal-metal interaction. The structural trans-influence at Tc and the network of H-bonds are consistent with Tc in oxidation state IV.  相似文献   

11.
The kinetics of the reaction between 1,4,8,11-tetraazacyclotetradecane (Cy) and Ni2+ in the presence of series of ligands L = fluoride, acetate, glycolate, oxalate, malonate, succinate, methanetriacetate, 1,3,5-cyclohexanetriacetate, tricarballylate, picolinate, glycinate, iminodiacetate, nitrilotriacetate. N,N′ -ethylenediiminodiacetate, ammonia, pyridine, ethylenediamine, 1,3-propanediamine and diethylenetriamine were studied by pH-static and spectrophotometric methods at 25° and I = 0.5. By analysis of the log k/log [L] and/or log k/pH profiles the resolved bimolecular rate constants K (Table 3) were determined using a non-linear least-square fitting procedure. Practically for all systems the rate constant K, describing the reaction between the 1:1 Ni2+ complex and the monoprotonated form of the macrocycle, was obtained. In some cases, however, also K and K were found. Since the experimental conditions were choosen so that NiL was mainly formed, the reactivity of NiL2 was generally not measurable. The effect of the number of coordinated donor groups in NiL and of the charge of NiL on K is discussed. Both effects seem to indicate that for the reaction between NiL and CyH+ first bond formation is not the rate-determining step.  相似文献   

12.
In aqueous acetonitrile (AN), Cu (I) forms the complexes Cu(AN)L+ and CuL with a series of substituted imidazoles (L). Stability constants logK of Cu(AN)+ + L ? Cu(AN)L+ and logβ2 were near 5 and 12, resp., log units for all ligands. The rate of autoxidation is described by ?d[O2]/dt=[CuL]2[O2](ka/(1+kb[CuL]) + (kc[L]+kd)/([CuL] + ke[Cu])), implying competition between one- or two-electron reduction of O2. The value of kc decreases from 5500M ?2S ?1 for unsubstituted imidazole to about 40M ?2S ?1 for 2-methylimidazole or 1,2-dimethyl-imidazole and essentially zero for the corresponding 2-ethyl-derivatives. On the other hand, ka and kb are much less influenced by the nature of the ligands, all values being near 5 · 104M ?2S ?1 and 103M ?1, respectively, for the complexes with the last four bases. Thus rather subtle sterical changes may strongly influence the relative importance of different pathways in the reduction of dioxygen by cuprous complexes.  相似文献   

13.
The stabilities of the Mn2+-, Co2+-, Ni2+-, Cu2+- and Zn2+-complexes with 2-(carboxymethyl)glutaric acid ( 2 ) and cis,cis-1,3,5-cyclohexanetricarboxylic acid ( 3 ) were measured potentiometrically at 25° and I = 0.5 (KNO3). Beside the complexes ML? protonated species MLH and MLH are also formed. Their stability constants are given in Table 1. A comparison between the stabilities of 2 or 3 and those of acetate, as a model for a monocarboxylate, or succinate and glutarate, as examples for dicarboxylates, indicates that in all species only one carboxylate is strongly bound whereas the second and third ones are probably not. The observation that Δlog K1 = log K ? log K as well as Δlog K2 = log K ? log K are practically constants with values of 0.34 ± 0.05 and 0.49 ± 0.07, respectively, for both ligands and the five metal ions studied is also in line with the proposed monodentate structures of the complexes ML?, MLH and MLH.  相似文献   

14.
The autoxidation of CuI in aqueous MeCN has been studied using a Clark oxygen electrode in the presence and absence of Cu11. The reaction is inhibited by Cu11 in the pH range of 0.5 to 5.0, reaching a lower limiting value at the highest concentrations. The reaction order changes from 1 to 2 with respect to CuI under the influence of Cu2+ ion. Detailed kinetics analysis of a total of 275 measurements has shown that an unstable primary adduct CuO+2 decomoses to give .O or HO, depnding on pH, and also reacts directly with a second Cu+ ion, avoiding one-electrton reduction of O2 by this path. Reaction of HO is faster with CuI than with Cu11 by a factor of 20, and single-electron transfer within CuO+2 to Cu2+ and .O predominates over reaction with a second copper ion for [CuItot] < 2. 10?3M in the absence of Cu2+. The most likely value for the reaction of .O with CuI is 5.3 · 108 M ?1S?1, but even this high rate constant is at the limit of significance. All secondary reactions followinfg the initial formation of CuO+2 are shown to be very fast, a fact that should be properly considered in the discussion of mechanisms of copper-catalyzed oxidations and oxygenations.  相似文献   

15.
The liquid phase fractionation factors αH = (18O/16O)H2O/(18O/16O)xo and αD = (18O/16O)D2O/(18O/16O)xo (X = Cl, Br, I) were calculated quantum mechanically between 0 and 100°. Experimental values were obtained in the case of BrO at 60° showing good agreement with the calculated results.  相似文献   

16.
The crystal structures of four anion cryptates [X? ? BT -6H+] formed by the protonated macrobicyclic receptor BT -6H+ with F?, Cl?, Br? and N have been determined. They provide a homogeneous series of anion coordination patterns with the same ligand. The small F?-ion is tetracoordinated, while Cl? and Br? are bound in an octahedron of H-bonds. The non-complementarity between these spherical anions and the ellipsoïdal cavity of BT -6H+ is reflected in ligand distortions. Structural complementarity is achieved for the linear triatomic substrate N, which is bound by two pyramidal arrays of three H-bonds, each interacting with a terminal N-atom of N. The formation constants of the complexes formed by BT -6H+ with a variety of anions (halides, N, NO, carboxylates, SO, HPO, AMP2?, ADP3?, ATP4?, P2O) have been determined. Very strong complexations are found, as well as marked electrostatic and structural effects on stability and selectivity; in particular the binding of F?, Cl?, Br?, and N may be analyzed in terms of the crystal structure data. The cryptand BT -6H+ is a molecular receptor containing an ellipsoïdal recognition site for linear triatomic substrates of size compatible with the size of the molecular cacity. Further developments of various aspects of anion coordination chemistry are considered.  相似文献   

17.
The complexation properties of the open-chain N2S2 ligands 1–4 are described and compared to those of analogous N2S2 macrocycles 5–7 . With Cu2+, the open-chain ligands give complexes with the stoichiometry CuL2+ and CuLOH+, the stabilities and absorption spectra of which have been determined. The ligand field exerted by these ligands is relatively constant and independent of the length of the chain. With Cu+, the species CuLH, CuLH2+, and CuL+ were identified and their stabilities measured. The redox potentials calculated from the equilibrium constants and measured by cyclic voltammetry agree and lie between 250 and 280 mV against SHE. The comparison between open-chain and cyclic ligands shows that (1) a macrocyclic effect is found for Cu2+ but not for Cu+, (2) the ligand-field strength is very different for the two types of ligands, and (3) the redox potentials span a larger interval for the macrocyclic than for the open-chain complexes.  相似文献   

18.
Nucleosides and Nucleotides, Part 11. Phosphorylation of 1-(2′-Desoxy-β-D-ribofuranosyl)-2(1H)-pyridon and its Behaviour in the Synthesis of Dinucleotides The behaviour of the unnatural nucleoside 1-(2′-deoxy-β-D-ribofuranosyl)-2(1H)-pyridon (Πd, 1 ) in the synthesis of dinucleotides with purine deoxynucleotides was studied. The optimized preparation of the protected dinucleoside phosphates (MeOTr) Πd pG ( 5 ) and (MeOTr) Πd pA ( 7 ) using the diester method of Khorana with DCC as condensing agent is described. The removal of the N-acyl- and p-methoxytrityl groups was effected by successive treatment with conc. ammonia solution and acetic acid/water 1:1 at 23° yielding the free dinucleoside phosphates ΠdpGd ( 9 ) and ΠdpAd ( 11 ). In a similar way, starting from (CNEt) pΠd( 15 ), the dinucleotides pΠdpG ( 16 ), pΠdpGd ( 18 ), pΠdpA ( 17 ) and pΠdpAd ( 19 ) were synthesized. The nucleotide 1-(5′-O-Phosphoryl-2′-deoxy-β-D-ribofuranosyl)-2(1H)-pyridon (pΠd, 3 ) was prepared in excellent yield by selective phosphorylation of Πd ( 1 ) using phosphorylchloride in triethyl phosphate at ?40°. Deoxyadenosine was phosphorylated in the same way. The compounds were characterized by UV. spectroscopy, chromatography and enzymatic degradation.  相似文献   

19.
The study at 25°C of the system K+? NH? CrO? SO? H2O has shown experimentally the existence of a new type of quaternary system of solubility with two cations and two anions. The solubility diagramm is caracterized by the presence of two adjacent ternary limiting systems with a miscibility gap, three univariant lines (one of them being evanescent), one invariant point, three binary and one ternary miscibility gaps.  相似文献   

20.
The 12-16 membered tetraazamacrocycles 1 - 6 were synthesized, their protonation constants and complexation kinetics measured at 25° and I = 0.50. The results of Table 1 Show that pK is strongly influenced by the ring size whereas pK and pK are relatively insensitive to it. This can be understood in terms of electrostatic interactions of the positive charges when located on adjacent amino groups. The kinetics of complex formation between the macrocyclic ligands and several transition metal ions have been studied by pH-stat and stopped-flow techniques and the results have been analyzed as bimolecular reactions between the metal ion and the different protonated species of the ligands. The rate constants, given in Table 2, show that the macrocycles react less rapidly than analogous open chain amines. However, for a given protonated species of the ligand the rate of complexation follows the order Cu2+ > Zn2+ > Co2+ > Ni2+ which parallels the sequence of their water exchange rates. For the diprotonated tetraamines LH reacting with Cu2+ the slower rates seem to be mainly a consequence of electrostatic interactions, since a correlation between logk and pK exists. For LH+, however, the complexation rates of a metal ion with the different macrocycles are all in one order of magnitude and do not depend in a regular way on the ring size or the basicity of the ligand. It is therefore suggested that in this case other factors such as unfavourable preequilibria must be considered as important.  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