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 15 毫秒
1.
Characteristics of the 13C-n.m.r. spectra of cellulose ethers (methyl, carboxymethyl, and hydroxyethyl) have been examined at 22.6 MHz. Partial depolymerization with acid or cellulase proved to be a requisite preliminary step. Strong deshielding of 13C nuclei bearing alkoxyl groups was clearly evident in these spectra, which permitted an assessment of the degree of substitution at individual positions of the d-glucose residues. Better resolved spectra, and more-detailed structural analyses, were afforded by complete hydrolysates of the polymers. The findings are wholly consistent with data obtained for these derivatives by other methods, showing that the reactivities of the hydroxyl groups of cellulose are OH-2>OH-6 ? OH-3. It is also shown that reducing-end residues liberated during enzymic hydrolysis of the cellulose derivatives are not substituted at the 2-position.  相似文献   

2.
The binding of cations of β-casein at pH 6.6 was considered previously. Available for three sodium concentiations, I = 0.04, 0.08, or 0.16 M are: [1] proton releases between I and [2] for each I, as calcium activity is increased, correlated sequences of monomer net charge, proton release, site bound calcium and protein Solvation- Models for ion binding are examined. Critical considerations are the intrinsic binding constants between hydrogen[H], calcium[Ca] and sodium[Na] ions and phosphate[P] and caiboxyIate[C] sites, and the effects of electrostatic interaction between sites as influenced by spatial fixed charge distribution, ionic strength and dielectric constant [D]. Anticipated intrinsic binding constants are kH,Po = 3 × 106, kCa,Po = 120, kNa,Po = 1, kH,Co = 7 × 104 and kCa,Co = 5.6Distributed charge models, either surface or volume, are inadequate since any reasonable monomer size yields fixed charge densities requiring kH,Po and kCa,Co which are too low when the maximum in D is 75. Also, with increasing calcium binding, calculated proton release is only 0.4 to 0.5 of that observed.Discrete charge models accept anticipated ko and yield calculated sequences of calcium binding and proton release which are in good agreement with those observed provided that: (1) using the known amino acid sequence of the phosphate-containing acidic peptide portion of the molecule, pep tide fixed charge is distributed at the lowest I so as to minimize electrostatic free energy; (2) in the region of fixed charge, D is approximately 5; (3) the distances between peptide fixed charges decrease with increasing ionic strength or calcium binding and (4) while protein is in solution, the acidic peptide and the remainder of the molecule are essentially electrostatically independent.  相似文献   

3.
《Bioorganic chemistry》1987,15(2):167-182
The kinetics of the Ni(II)-catalyzed ester hydrolysis of O-acetyl-2-pyridine-carboxaldoxime, O-acetyl-2-acetylpyridineketoxime, and O-acetyl-6-carboxy-2-pyridine-carboxaldoxime are measured and the values of various kcat parameters are calculated for reaction paths involving one metal ion (kcatW and kcatOH) and two metal ions kcatA and kcatB). Examination of the kinetic data reveals that the kcatW and kcatOH paths for the Ni(II)-catalyzed reactions involve the same mechanism as those for the previously reported Cu(II)-catalyzed reactions. For the kcatA and kcatB paths, the mechanism involving binuclear Ni(II) ions is preferred by analogy with the previously reported Zn(II)-catalyzed reactions. Comparison of kcatOH values for the Cu(II)- and Ni(II)-catalyzed hydrolysis of 1–3 indicates that markedly different steric effects are exerted by the substituents of 2 and 3 on the catalytic behavior of the two metal ions. This is explained in terms of differences in the fit of the metal ions in the metal complexes of 1–3. Present results demonstrate that slight changes in the geometry around the central metal atom can affect the catalytic outcome significantly. The implications of the present results on metal substitution in metalloenzymes are also discussed.  相似文献   

4.
Mononuclear nonheme iron enzymes (MNHEs) catalyze a range of very diverse reactions in O2 metabolism, but they share a common principle active-site organization. To investigate a putative catalytic promiscuity of these enzymatic metal centers, we studied the reactivity of the 3-His ligated metal center of diketone cleaving enzyme (Dke1) toward non-native substrates, with a focus on alternative O2 dependent reactions. From a screening approach, which aims at eliminating steric factors by including minimal substrate-substructures, three alternative, ‘non-β-dicarbonyl-cleavage’ reactions are identified, among them an unprecedented oxygenation of maltol. Maltol cleavage is characterized by steady state and fast kinetic measurements and shows an O2 concentration dependent rate determining step kcat/KM(O2) of 0.3 mM− 1 s− 1 and a strict coupling of O2 reduction and substrate oxidation. Furthermore, the catalytic potential of the 3-His metal center for O2 dependent catechol ring-cleavage and phenylpyruvate oxidation (PP) is demonstrated.  相似文献   

5.
The oxygen isotope ratios of tree ring cellulose have a great potential as proxy for the oxygen isotope ratios of source water, which is related to climate. However, source water isotopic signatures can be masked by plant physiological and biochemical processes during cellulose synthesis. To minimize biochemical effects in the recording of source water, we modified the cellulose molecule to phenylglucosazone, which only has oxygen attached to carbon 3–6 (OC3–6) of the cellulose glucose moieties, thus eliminating the oxygen attached to carbon 2 of the cellulose glucose moieties (OC-2). Here we developed a method to use small amounts of inter and intra-annual tree ring cellulose for phenylglucosazone synthesis. Using this new method we tested if the oxygen isotope ratios of source water reconstructed from tree ring phenylglucosazone (δ18OswPG) and the observed source water (δ18Oswobs) would have a better agreement than those reconstructed from the tree ring cellulose molecule. Annual tree ring samples were obtained from Pinus sylvestris (1997–2003) (Finland) and Picea abies (1971–1992) (Switzerland) and intra-annual tree ring samples were obtained from Pinus radiata (October 2004–March 2006) (New Zealand), each near a meteorological station where precipitation and relative humidity (RH) were measured periodically. The δ18O of tree ring cellulose and tree ring phenylglucosazone for each of the three species were then used to back calculate the δ18O of source water according to a previous published empirical equation. As expected, the δ18O of tree ring phenylglucosazone was superior than cellulose in the reconstruction of source water available to the plant. Deviation between δ18OswPG and δ18Oswobs was in part correlated with variation in atmospheric relative humidity (RH) which was not observed for the cellulose molecule. We conclude that this new method can be applicable to inter and intra-annual tree ring studies and that the use of the tree ring phenylglucosazone will significantly improve the quality of paleoclimate studies.  相似文献   

6.
Three methods of wheat straw treatment (with NaOH, H2O2 and butylamine) and its subsequent saccharification by Trichoderma reesei cellulases and Aspergillus niger β-glucosidase are reported. The treatment of straw with NaOH for 1 h in the autoclave (120°C) caused a great loss of the hemicellulose content and a partial removal of lignin, provoking an increase of the cellulose content (from 24% to 69%) in the residue. When the straw was pre-treated with H2O2 at 25°C for 20 h, the relative content of cellulose in the straw increased (from 24% to 52%) due to the solubilisation of hemicellulose.

The effect of varying the hydrolysis parameters (enzyme and substrate concentration, temperature and pH) was studied in order to maximise the yield of sugars. Under the best conditions and after 48 h with a mixture of 2:1 (w/w) cellulase/β-glucosidase (with a concentration of 7 and 0.1 U ml-1, respectively) the native, NaOH-treated and H2O2-treated straw material were degraded to reducing sugars for 28%, 89% and 97% respectively.  相似文献   

7.
The rates and mechanisms for the reactions between Trolox (6-hydroxy-2,5,7,8-tetramethylchroman-2-carboxylic acid), a synthetic analogue of α-tocopherol, Br1 and Mn3+-phosphate were determined under aerobic and anaerobic conditions by fast kinetic methods. In addition, the reaction between Fe(OH)2+ and Trolox at pH 6.0 was shown not to occur under the conditions of the experiment, leading to a limiting rate for that reaction of k < 103M−1s−1. These results are compared to studies of the reactivity of Trolox with HO2/O2 and are discussed in light of the known antioxidant properties of vitamin E.  相似文献   

8.
In a study of the isolobal analogy between the proton H+ and the ligand-backed gold(I) cations [(R3P)Au]+, the reaction of the mixture of 2-pyridone/2-hydroxy-pyridine tautomers with [(Ph3P)Au]BF4 has been investigated. It affords the 1:1 complex with the gold atom N-bonded to the 2-hydroxy-pyridine tautomer: {(Ph3P)Au[NC5H4-(OH-2)]}+BF4 −, which is related to known salts with the 2-hydroxy-pyridinium cation such as [HNC5H4(OH-2)]+Cl. The structure was derived from analytical and spectroscopic data and from a comparison with the salt [(Ph3P)Au(pyr)]BF4, prepared and investigated structurally as a reference compound. An analogue was also prepared with 2-dimethylamino-ethanol as a substrate, which also affords the N-bonded complex [(Ph3P)Au([Me2NCH2CH2OH)]+BF4 −, the structure of which has been determined. The OH group is not attached to the gold atom but engaged in hydrogen bonding with the counterion. By contrast, in the complex [(Ph3P)Au(Me2NCH2CH2NMe2)]+BF4 − synthesized similarly with tmeda and crystallized as the dichloromethane solvate, one nitrogen atom is bonded firmly to the metal atom, but the second nitrogen atom is also weakly engaged in coordinative bonding. The compound is fluxional in solution, where a site exchange is observed which is rapid on the NMR time scale. The reaction of two equivalents of [(Ph3P)Au]BF4 with an alkali 2-pyridinolate, prepared from the above tautomeric mixture and sodium metal or a potassium alkoxide, yields the diaurated product {N,O-[(Ph3P)Au]2(NC5H4-O-2)}BF4. In the crystal structure determination of a sesqui-solvate with dichloromethane it has been shown that one gold atom is attached solely to the nitrogen atom and the other solely to the oxygen atom. The dinuclear cations are associated into cyclic dimers through head-to-tail aurophilic bonding. From the geometrical characteristics of the core unit of the cations the ligand can be assigned a 2-pyridinolate form featuring pyridine and phenolate donor sites.  相似文献   

9.
The nonbonded interaction energy of disaccharides, mannobiose and galactobiose and polysaccharides mannan and galactan have been computed as a function of dihedral angles (?,ψ). The conformation (40°, ?20°) has been preferred for the mannan chain from nonbonded interaction energy considerations. The O5…O3′ type of intramolecular hydrogen bond has been found to be possible in the above conformation. Comparison of the allowed region of mannan with those of cellulose and xylan indicates that the monomer unit, in mannan chain has slightly higher freedom of rotation than that of cellulose and less than that of xylan. As in cellulose and mannan, the freedom of rotation of the monomer units in β-1,4′ galactan is highly restricted. Unlike mannan (which prefers an extended conformation) the β-1,4′ galactan prefers a helical conformation similar to amylose. Just as in amylose the O2…O3′ type hydrogen bond between contiguous residues is also possible in β-1,4′ galactan.  相似文献   

10.
Available data on the kinetic processes in H2-O2-O2(a 1Δ g ) mixtures are analyzed theoretically, and the ranges in which the rate constants of these processes can vary are determined. The processes of energy transformation in an O2(a 1Δ g )-H2-H-HO2 system in the approximations of the fast and slow (in comparison with the vibrational relaxation time of the HO2 radical) quenching of the electronically excited state are considered. The experiments on the quenching of singlet delta oxygen (SDO) molecules O2(a 1Δ g ) excited in a microwave discharge at a temperature of 300 K and pressure of 6 Torr in a lean hydrogen-oxygen mixture are simulated (by a lean fuel mixture is meant a mixture in which the ratio of the fuel to the oxidizer mass fraction is less than the stoichiometric ratio, which is 2: 1 for a hydrogen-oxygen mixture). It is shown that, over the experimental observation times, the SDO quenching frequency depends on the concentration of molecular hydrogen, the residual odd oxygen fraction in the gas flow, and the ratio between the rate constants of kinetic processes involving HO2 and HO2* radicals. Simulations of the microwave discharge and the transport of excited oxygen along the drift tube make it possible to determine the residual odd oxygen concentration in the gas flow. Recommendations on the choice of the rate constants for the O2(a 1Δ g ) + HO2)A″, v3″ = 0) ? O2 + HO2*(A′, v3′ = 1), HO2*(A′v3′ ≤ 1) + O2(a 1Δ g ) → HO2*(A′,v3′ ≥ 6) + O2, and HO2*(A′,v3′ ≤ 1) + O2(a 1Δ g ) → H + O2 + O2 processes are given. It is shown that, in the case of slow quenching in a H2-O2-O2(a 1Δ g ) mixture at a low temperature, the intensity of hydrogen oxidation is enhanced due to the reaction + HO2*(A′,v3′ ≤ 1) + O2(1Δ) → H + O2 + O2.  相似文献   

11.
Comb-shaped copolymers with cellobiose acetate or cellulose triacetate (CTA) side-chains, PPMA-g-(CTA2-C15) and PPMA-g-(CTA13-C15), were prepared by grafting N-(15-azidopentadecanoyl)-2,3,6-tri-O-acetyl-4-O-(2,3,4,6-tetra-O-acetyl-β-d-glucopyranosyl)-β-d-glucopyranosylamine (CTA2-C15-N3) and N-(15-azidopentadecanoyl)-tri-O-acetyl-β-cellulosylamine (CTA13-C15-N3, number average degree of polymerization (DPn= 13) onto poly(2-propyn-1-yl methacrylate) (PPMA, weight average degree of polymerization (DPw, X + Y = 5.59 × 102)) via “click chemistry”. The copolymers were characterized by 1H, 13C and two-dimensional NMR and size exclusion chromatography-multi-angle laser light scattering (SEC-MALS) measurements. The numbers of CTA side-chains (X) of PPMA-g-(CTA2-C15) and PPMA-g-(CTA13-C15) were calculated as 4.03 × 102 and 2.45 × 102, respectively. Copolymers with cellulosic side-chains, PPMA-g-(CELL2-C15) and PPMA-g-(CELL13-C15), were successfully obtained after deacetylation of PPMA-g-(CTA2-C15) and PPMA-g-(CTA13-C15), respectively. X-ray diffraction measurements revealed that PPMA-g-(CELL13-C15) showed crystalline pattern of cellulose II, which is believed to have anti-parallel orientation.  相似文献   

12.
A mononuclear cobalt(III)-peroxo complex bearing a macrocyclic tetradentate N4 ligand, [CoIII(TMC)(O2)]+ (TMC = 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane), was generated in the reaction of [CoII(TMC)]2+ and H2O2 in the presence of triethylamine in CH3CN. The reactivity of the cobalt(III)-peroxo complex was investigated in aldehyde deformylation with various aldehydes and compared with that of iron(III)- and manganese(III)-peroxo complexes, such as [FeIII(TMC)(O2)]+ and [MnIII(TMC)(O2)]+. In this reactivity comparison, the reactivities of metal-peroxo species were found to be in the order of [MnIII(TMC)(O2)]+ > [CoIII(TMC)(O2)]+ > [FeIII(TMC)(O2)]+. A positive Hammett ρ value of 1.8, obtained in the reactions of [CoIII(TMC)(O2)]+ and para-substituted benzaldehydes, demonstrates that the aldehyde deformylation by the cobalt(III)-peroxo species occurs via a nucleophilic reaction.  相似文献   

13.
Condensation of 6-O-benzyl-7,8-dideoxy-1,2:3,4-di-O-isopropylidene-d-glycero-α-d-galacto-oct-7-ynopyranose with methyl 2,3,4-tri-O-benzyl-6-deoxy-β-d-galacto-heptodialdo-1,5-pyranoside afforded a 2:1 mixture of the 1S and 1R isomers (1a and 1b) of 3-[6(R)-O-benzyl-1,2:3,4-di-O-isopropylidene-α-d-galactopyranos-6-yl]-1-hydroxy-1-(methyl 2,3,4-tri-O-benzyl-6-deoxy-β-d-galactopyranosid-6-yl)propyne. A single crystal of the 1-O-acetyl derivative (1c) of 1a was investigated by X-ray diffraction methods in a four-circle diffractometer. Compound 1c crystallises in the monoclinic system, space group P21 (Z = 2) with cell dimensions a = 14.896(2), b = 8.295(1), c = 20.547(3) Å, and β = 102.66(1)°. The structure was solved by direct methods and refined by a full-matrix, least-squares procedure against 3839 unique reflections (F > 2σF), resulting in a final R = 0.045 (unit weights). The configuration at the new chiral center (C-1) was established as S(d). The galactopyranose rings have conformations 4C1 (tri-O-benzylated moiety) and °S5 + °T2 (di-O-isopropylidenated moiety). The 1,2- and 3,4-O-isopropylidene rings have 3T2 and 2E conformations, respectively.  相似文献   

14.
trans-[Ru(NH3)4P(OR)3(H2O)]2+ (R = Me, Pr, iPr, and Bu) reacts with isonicotinamide at second-order- specific rates k1 of 1.2, 2.3, 7.4 and 8.1 M−1 s(25 °C, μ = 0.10 NaCF3COO/CH3COOH), respectively, for R = Me, Pr, iPr and Bu. The products trans- [Ru(NH3)4P(OR)3isn](PF6)2 have been isolated and characterized by micro analysis, cyclic voltammetry, and electronic spectral data. The aquation rates k−1 for the isonicotinamide (isn) derivatives are 5.2 × 10−2, 5.9 × 10−2, 2.0 × 10−1 and 3.4 × 10−1 s−1 for R= Me, Pf, Bu and iPr, respectively. The activation parameters for the forward and backward reactions indicate the same mechanism for all of them. The substitution proceeds by a dissociative mechanism with a significant outer-sphere association of trans-[Ru(NH3)4P(OR)3(H2O)]2+ complexes with isn. Assuming k1 as indicative of the lability of the coordinated water molecule on the monophosphite complexes, the following sequence of increasing trans-effect mav be proposed: P(OMe)3 <P(OEt)3 <P(OPr)3 <P(OiPr)3 <P(OBu)3. The affinity of the monophosphite complexes for isn increases according to P(OMe)3 ⋍ P(OiPr)3 < P(OEt)3 < P(OPr)3 ⋍ P(OBu)3.  相似文献   

15.
《Inorganica chimica acta》1988,141(2):211-220
The reaction of CrCl3 · 6H2O (dehydrated in DMSO) with 1,5,9-triazanonane (3,3-tri) gives mer- CrCl3(3,3-tri), the configuration being established by isomorphism with the corresponding Co(III) complex. This non-electrolyte is hydrolyzed in aqueous acidic solution and mer-[CrCl2(3,3-tri)- (OH2)]ClO4 can be isolated by anation with HCl in the presence of HClO4. Reaction of mer-CrCl3- (3,3-tri) in DMF with diamines produces complexes of the type [CrCl(diamine)(3,3-tri)] Cl2 [diamine= 1,2-diaminoethane (en), 1.2-diaminopropane (pn), 1,3-diaminopropane (tn), 2,2-dimethyl-1,3-diaminopropane (Me2tn) and cyclohexanediamine (chxn, cis plus trans mixture; two isomers A and B)] and these have been characterized as the ZnCl42− salts. The configuration of the triamine ligand in these complexes has been established as mer-(H↓)- by a single crystal X-ray analysis of [CrCl(en)(3,3-tri)]- ZnCl4, monoclinic, P21, a=7.932, b= 14.711, c= 8.312 Å, β=104.6° and Z=2, refined to a conventional R factor of 0.034. The kinetics of the Hg2+- assisted chloride release from [CrCl(diamine)(3,3- tri)]ZnCl4 salts were measured spectrophotometrically (μ=1.0 M HClO4 or HNO3) over 15 K temperature ranges to give, in order, 104kHg (298.2 K) (M−1 s−1), Ea(kJ mol−1), ΔS# (J K−1 mol−1): en- (HClO4): 5.95, 78.1, -53; pn(HClO4); 5.24, 81.2; -44; tn(HClO4): 26.7, 85.6, -15; Me2tn(HClO4): 21.8, 78.6, -40; A-chxn(HNO3): 7.60, 81.0,-41; B-chxn(HNO3): 18.3, 56.8, -115. A ‘non-replaced ligand effect’ on the rate is observed for the first time in this series of homologous Cr(III) complexes. The kinetics of the thermal aquation (kH, 0.1 M HClO4) were measured titrimetrically for CrCl(diamine) (3,3-tri)2+ to give the following kinetic parameters: diamine=en: 107 kH (298.2)=5.34 s−1, Ea=99.2 kJ mol−1, ΔS#=-40 J K−1 mol-1; diamine =tn: 107 kH (298.2)=5.04 s−1, Ea= 82.8, ΔS#= -96.  相似文献   

16.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  相似文献   

17.
The rate of Fe3+ release from horse spleen ferritin (HoSF) was measured using the Fe3+-specific chelator desferoxamine (DES). The reaction consists of two kinetic phases. The first is a rapid non-linear reaction followed by a slower linear reaction. The overall two-phase reaction was resolved into three kinetic events: 1) a rapid first-order reaction in HoSF (k1); 2) a second slower first-order reaction in HoSF (k2); and 3) a zero-order slow reaction in HoSF (k3). The zero-order reaction was independent of DES concentration. The two first-order reactions had a near zero-order dependence on DES concentration and were independent of pH from 6.8 to 8.2. The two first-order reactions accounted for 6-9 rapidly reacting Fe3+ ions. Activation energies of 10.5 ± 0.8, 13.5 ± 2.0 and 62.4 ± 2.1 kJ/mol were calculated for the kinetic events associated with k1, k2, and k3, respectively. Iron release occurs by: 1) a slow zero-order rate-limiting reaction governed by k3 and corresponding to the dissociation of Fe3+ ions from the FeOOH core that bind to an Fe3+ binding site designated as site 1 (proposed to be within the 3-fold channel); 2) transfer of Fe3+ from site 1 to site 2 (a second binding site in the 3-fold channel) (k2); and 3) rapid iron loss from site 2 to DES (k1).  相似文献   

18.
The magnesium ion-dependent equilibrium of vacant ribosome couples with their subunits
70 S?k?1k150 S+30S
has been studied quantitatively with a novel equilibrium displacement labeling method which is more sensitive and precise than light-scattering. At a concentration of 10?7m, tight couples (ribosomes most active in protein synthesis) dissociate between 1 and 3 mm-Mg2+ at 37 °C with a 50% point at 1.9 mm. The corresponding association constants Ka′ are 5.1 × 105m?1 (1 mm-Mg2+), 3.5 × 107m?1 (2 mm), and 1.2 × 109m?1 (3 mm), about five orders of magnitude higher than the Ka′ value of loose couples studied by Spirin et al. (1971) and Zitomer & Flaks (1972).In this range of Mg2+ concentrations (37 °C, 50 mm-NH4+) the rate constants depend exponentially and in opposite ways on the Mg2+ concentration: k1 = 2.2 × 10?3s?1, k?1 = 7.7 × 104m?1s?1 (2mm-Mg2+); k1 = 1.5 × 10?4s?1, k?1 = 1.7 × 107m?1s?1 (5 mm-Mg2+). Under physiological conditions (Mg2+ ~- 4 mm, ribosome concn ~- 10?7m), the equilibrium strongly favors association and the rate of exchange is slow (t12 ~- 10 min). In the range of dissociation (2 mm-Mg2+), association of subunits proceeds without measurable entropy change and hence ΔGO = ΔHO. The negative enthalpy change of ΔHO = ? 10 kcal suggests that association of subunits involves a shape change.Below a critical Mg2+ concentration (~- 2 mm), the 50 S subunits are converted irreversibly into the b-form responsible for the transition to loose couples. The results are compatible with two classes of binding sites, one class binding Mg2+ non-co-operatively and contributing to the free energy of association by reduction of electrostatic repulsion, and another class probably consisting of hydrogen bonds between components at opposite interfaces whose critical spatial alignment rapidly denatures in the absence of stabilizing magnesium ions.  相似文献   

19.
Several platinum(II) complexes of the general type [Pt(OND)X] have been prepared and characterized, the ligand (OND) representing the phenolate anion of the tridentate Schiff bases N-(2- diethylaminoethyl)-salicylaldimine (D = NEt2), N-(2- ethylaminoethyl)-salicylaldimine (D = NHEt) and N- (3-thia-n-pentyl)-salicylaldimine (D= SEt) and X= Cl, NO3. As shown by conductimetric studies the nitrato complexes [Pt(OND)NO3] dissociate completely in methanol according to:
Spectrophotometry (normal and stopped-flow) has been used to study the kinetics of solvent substitution according to
with a variety of neutral and anionic nucleophiles Y in methanol at 20 °C and constant ionic strength, I= 0.2 M (NaClO4). The substitution follows a one- term rate law, v = kobs[Pt(OND)(H2O)+] = kY[Y]- [Pt(OND)(H2O)+]. The kY data obtained for 13 (D = NEt2) and 7 (D = NHEt; SEt) different nucleophiles Y cannot be adequately correlated with their npt0 values according to the well-known relationship log kY = snpt0 + log ks. The deviations are strongest for large and bulky nucleophiles such as Y=Ph3P, Bu3P, Ph3As, I- and for D = NEt2, from which it is concluded that steric crowding hinders the formation of the 5-coordinate transition state. The rate reducing steric cis-effect observed is of the order kY(D = NEt2):kY(D = NHEt):kY(D = SEt) = 1:35:63 for small nucleophiles Y and as large as 1:192:2640 for Y = Ph3P. The introduction of substituents X in the salicylaldehyde ring in ortho (X3), meta (X4) and para position (X5) to the phenolic oxygen proves the existence of rather small electronic effects (X4, X5) and much stronger steric effects of bulky substituents X3, neighboring the donor oxygen.With the standard substrate trans-[Ptpy2Cl2] some new npt0 values were determined, namely for N, N′- dimethylthiourea (npt0 = 7.02), N, N′ -diphenylthlourea (npt0 = 7.19), N, N, N′, N′-tetramethylthiourea (npt0 = 6.05) and for the pseudo-halide dicyanoamide ion, N(CN)2- (npt0= 3.05). The npt0 value for the pseudo-halide tricyanomethanide, ion, C(CN)3-, was estimated to be 3.03.  相似文献   

20.
Measurements of the equilibrium and temperature-jump u.v., visible, and induced c.d. spectra of Methyl Orange (MO) in the presence of cyclomalto-octaose (γ-cyclodextrin, γ-CD) have been carried out. Three mechanistic steps were detected through the temperature-jump data (25.0°):
where K1, K2, and K3 are 45 (±7), 2.0 (±1.1) × 106, and 6.1 (±2.5) × 103 dm3.mol?1, respectively, k2 = 9.4 (±5.1) × 109 dm3.mol?1.s?1, and k?2 = 4.8 (±0.8) × 103 s?1. The equilibrium u.v./visible data are also consistent with this reaction scheme. The high stability of the dimer inclusion complex (MO)2 · γ-CD compared to that of the monomer inclusion complex MO · γ-CD appears to be related to the annular diameter of γ-CD and demonstrates a degree of selectivity in cyclodextrin inclusion complexes. The (MO)2 · (γ-CD)2 complex also contains a dimer, included by both γ-CD molecules.  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