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 46 毫秒
1.
Copper(II) complexes supported by bulky tridentate ligands L1H (N,N-bis(2-quinolylmethyl)-2-phenylethylamine) and L1Ph (N,N-bis(2-quinolylmethyl)-2,2-diphenylethylamine) have been prepared and their crystal structures as well as some physicochemical properties have been explored. Each complex exhibits a square pyramidal structure containing a coordinated solvent molecule at an equatorial position and a weakly coordinated counter anion (or water) at an axial position. The copper(II) complexes reacted readily with H2O2 at a low temperature to give mononuclear hydroperoxo copper(II) complexes. Kinetics and DFT studies have suggested that, in the initial stage of the reaction, deprotonated hydrogen peroxide attacks the cupric ion, presumably at the axial position, to give a hydroperoxo copper(II) complex retaining the coordinated solvent molecule (H R ·S). H R ·S then loses the solvent to give a tetragonal copper(II)-hydroperoxo complex (H R ), in which the –OOH group may occupy an equatorial position. The copper(II)–hydroperoxo complex H R exhibits a relatively high O–O bond stretching vibration at 900 cm−1 compared to other previously reported examples.Electronic Supplementary Material Supplementary material is available for this article at  相似文献   

2.
A laboratory study investigated the metabolic physiology, and response to variable periods of water and sodium supply, of two arid-zone rodents, the house mouse (Mus domesticus) and the Lakeland Downs short-tailed mouse (Leggadina lakedownensis) under controlled conditions. Fractional water fluxes for M. domesticus (24 ± 0.8%) were significantly higher than those of L. lakedownensis (17 ± 0.7%) when provided with food ad libitum. In addition, the amount of water produced by M. domesticus and by L. lakedownensis from metabolic processes (1.3 ± 0.4 ml · day−1 and 1.2 ± 0.4 ml · day−1, respectively) was insufficient to provide them with their minimum water requirement (1.4 ± 0.2 ml · day−1 and 2.0 ± 0.3 ml · day−1, respectively). For both species of rodent, evaporative water loss was lowest at 25 °C, but remained significantly higher in M. domesticus (1.1 ± 0.1 mg H2O · g−0.122 · h−1) than in L. lakedownensis (0.6 ± 0.1 mg H2O · g−0.122 · h−1). When deprived of drinking water, mice of both species initially lost body mass, but regained it within 18 days following an increase in the amount of seed consumed. Both species were capable of drinking water of variable saline concentrations up to 1 mol · l−1, and compensated for the increased sodium in the water by excreting more urine to remove the sodium. Basal metabolic rate was significantly higher in M. domesticus (3.3 ± 0.2 mg O2 · g−0.75 · h−1) than in L. lakedownensis (2.5 ± 0.1 mg O2 · g−0.75 · h−1). The study provides good evidence that water flux differences between M. domesticus and L. lakedownensis in the field are due to a requirement for more water in M. domesticus to meet their physiological and metabolic demands. Sodium fluxes were lower than those observed in free-ranging mice, whose relatively high sodium fluxes may reflect sodium associated with available food. Accepted: 16 August 1999  相似文献   

3.
 The stability constants of the 1 : 1 complexes formed between Mg2+ and the anions of the N1, N3, and N7 deaza derivatives of 9-[2-(phosphonomethoxy)ethyl]adenine (PA2–), i.e., of Mg(H;PA)+ and Mg(PA), were determined by potentiometric pH titration in aqueous solution (25  °C; I=0.1 M, NaNO3) and compared with previous results [Sigel H, et al. (1992) Helv Chim Acta 75 : 2634–2656], obtained under the same conditions, for the corresponding complexes of 9-[2-(phosphonomethoxy)ethyl]adenine (PMEA2–) and (phosphonomethoxy)ethane (PME2–). Based on the analysis of a microconstant scheme it is concluded that in the monoprotonated complexes, Mg(H;PA)+, Mg2+ is coordinated to a significant part at the nucleobase, H+ being at the phosphonate group. By making use of log K Mg Mg(R-PO3) versus pK H H(R-PO3) straight-line plots (also obtained previously; see above) for simple phosphonates and phosphate monoesters, it is shown that all the Mg(PA) complexes, including those with PMEA2– and PME2–, are more stable than expected on the basis of the basicity of the ―PO2– 3 group. This proves that, to some extent, five-membered chelates, Mg(PA)cl/O, involving the ether oxygen of the ―CH2―O―CH2―PO2– 3 chain are formed; their formation degree amounts to about 30–40% in equilibrium with the isomer having only a phosphonate-Mg2+ coordination. In the case of Mg(1-deaza-PMEA), probably a further isomer occurs in which also N3 of the nucleobase participates. The different properties between the Mg(PA) species and the Mg(AMP) complex are discussed. Received: 26 January 1998 / Accepted: 19 May 1998  相似文献   

4.
 The Gd(III) complex of 4-pentylbicyclo[2.2.2]octane-1-carboxyl-di-l-aspartyl-lysine-derived DTPA, [GdL(H2O)]2–, binds to serum albumin in vivo, through hydrophobic interaction. A variable temperature 17O NMR, EPR, and Nuclear Magnetic Relaxation Dispersion (NMRD) study resulted in a water exchange rate of k 298 ex=4.2×106 s–1, and let us conclude that the GdL complex is identical to [Gd(DTPA)(H2O)]2– in respect to water exchange and electronic relaxation. The effect of albumin binding on the water exchange rate has been directly evaluated by 17O NMR. Contrary to expectations, the water exchange rate on GdL does not decrease considerably when bound to bovine serum albumin (BSA); the lowest limit can be given as k ex, GdL-BSA=k ex, GdL / 2. In the knowledge of the water exchange rate for the BSA-bound GdL complex, the analysis of its NMRD profile at 35  °C yielded a rotational correlation time of 1.0 ns, one order of magnitude shorter than that of the whole protein. This value is supported by the longitudinal 17O relaxation rates. This indicates a remarkable internal flexibility, probably due to the relatively large distance between the protein- and metal-binding moieties of the ligand. Received: 25 June 1998 / Accepted: 11 August 1998  相似文献   

5.
The vertebrate renin-angiotensin system controls cardiovascular, renal and osmoregulatory functions. Angiotensin II (ANG II) is the most potent hormone of the RAS but in some vertebrate animals angiotensin III (Val4-ANG III) may be a hormone. We studied the effects of some angiotensins and mammalian ANG II receptor antagonists on nasal salt gland function and arterial blood pressure in conscious white Pekin ducks. Nasal salt gland fluid secretion (NFS) was induced by a 10 ml · kg−1 bw i.v. injection of a NaCl solution (1000 mosmol · kg−1 H2O) and maintained by a continuous i.v. infusion of the same solution at a rate of 0.97 ml · min−1. There was a positive linear correlation between nasal fluid [Na+] and osmolality, between [Na+] and [K+], and also between the rate of NFS and [Na+] and [K+]. [Asp1,Val5]-ANG II (1 nmol · kg−1 i.v.) inhibited NFS but did not change ionic concentrations. Val4-ANG III (1 or 5 nmol · kg−1) and ANG I (1-7) (20 nmol · kg−1) had no effect on NFS. [Sar1, Ile8]-ANG II (SARILE) acted as an ANG II receptor agonist and resulted in a prolonged and complete inhibition of NFS. The AT1 receptor antagonist, losartan (DuP 753) and the AT2 receptor antagonist, PD 123319 both failed to block the inhibitory effect of [Asp1, Val5]-ANG II on the nasal salt glands. [Asp1,Val5]-ANG II (2 nmol · kg−1 i.v.) increased mean arterial blood pressure (MABP), whereas the same dose of [Asn1,Val5]-ANG II (teleost) had only 30% of the pressor potency of the avian ANG II. Neither 1 nor 5 nmol · kg−1 of Val4-ANG III i.v. nor 20 nmol · kg−1 of ANG I (1-7) had any measurable effect on MABP. SARILE blocked completely the pressor response to [Asp1,Val5]-ANG II but the AT1 antagonists losartan and CGP 48933 and the AT2 antagonist PD 123319 all failed to block the pressor response to [Asp1,Val5]-ANG II. These results have substantiated an important role of the nasal salt gland in potassium regulation and highlighted a pharmacological dimorphism of saralasin, namely agonist and antagonist to angiotensin II-mediated inhibition of nasal salt gland function and pressor response, respectively. Using specific nonpeptidergic angiotensin II receptor antagonists, we have confirmed the distinct pharmacology of the avian angiotensin II receptors in a nongallinaceous species and the absence of significant angiotensin I (1-7) and angiotensin II effects on the cardiovascular system and nasal salt gland. Accepted: 6 November 1997  相似文献   

6.
The microsolvation of taurine (TA) with one, two or three water molecules was investigated by a density functional theory (DFT) approach. Quantum theory of atoms in molecules (QTAIM) analyses were employed to elucidate the hydrogen bond (H-bond) interaction characteristics in TA-(H2O)n (n = 1–3) complexes. The results showed that the intramolecular H-bond formed between the hydroxyl and the N atom of TA are retained in most TA-(H2O)n (n = 1–3) complexes, and are strengthened via cooperative effects among multiple H-bonds from n = 1–3. A trend of proton transformation exists from the hydroxyl to the N atom, which finally results in the cleavage of the origin intramolecular H-bond and the formation of a new intramolecular H-bond between the amino and the O atom of TA. Therefore, the most stable TA-(H2O)3 complex becomes a zwitterionic complex rather than a neutral type. A many-body interaction analysis showed that the major contributors to the binding energies for complexes are the two-body energies, while three-body energies and relaxation energies make significant contributions to the binding energies for some complexes, whereas the four-body energies are too small to be significant.  相似文献   

7.
A novel one-dimensional heterometallic complex, {Cd2[NiL]2(SCN)4(H2O)}n (1), has been synthesized and characterized by single-crystal X-ray analysis, where L is dianion of 2,3-dioxo-5,6,13,14-dibenzo-9,10-cyclohexyl-7,12-bis(ethoxycarbonyl)-1,4,8,11-tetraazacyclotetradeca-7,11-diene. The most striking feature of 1 is that in the structure there is one type of S-S bond (1.823(13) Å) formed by two thiocyanate groups which has not been reported to our knowledge. The DNA cleavage activity of 1 in the presence of H2O2 was compared with those of nickel(II) ion, cadmium(II) ion and corresponding mononuclear precursor NiL (2). The DNA cleavage kinetics was studied and the corresponding activation parameters of 1 were obtained.  相似文献   

8.
 The coordination state of Fe(III)- and Fe(II)-mimochrome I, a covalent peptide-deuteroheme sandwich involving two nonapeptides bearing a histidine residue in a central position, was studied by UV-visible, EPR, and resonance Raman spectroscopy. The ferric and ferrous states of this new iron species mainly exist, at pH 7, in a low-spin hexacoordinate form with two axial histidine ligands coming from the peptide chains. A minor amount of high-spin form for the ferric state is also present at pH 7. However, it is mainly high-spin at pH 2 or in DMSO. Fe(II)-mimochrome I binds CO with an affinity comparable to that of myoglobin and hemoglobin. Fe(III)-mimochrome I reacts with alkylhydroxylamine and arylhydrazines, leading to the corresponding Fe(II)-nitrosoalkyl and Fe(III)-σ-aryl complexes, respectively. These reactions were greatly dependent on the solvent used and on the pH, and were much slower than the corresponding reactions performed by deuterohemin in the presence of excess imidazole. All these results indicate that the reactivity of iron-mimochrome I is controlled by the binding of the peptide chains to the iron. The reactivity shown by this complex at neutral pH is intermediate between that observed for iron porphyrins in the presence of excess imidazole and that of hemoproteins characterized by a strong bis-histidine axial coordination, such as cytochrome b 5. Fe(III)-mimochrome I is able to catalyze styrene epoxidation by using a [Fe(III)-mimochrome I]/[H2O2]/[stryrene] ratio of 1 : 10 : 2000 in phosphate buffer solution (pH 7.2) containing 2% CTAB both under strictly anaerobic conditions and in the presence of oxygen, at 0  °C. Received: 26 May 1998 / Accepted: 20 August 1998  相似文献   

9.
 Kinetics of the steady-state oxidation of n–alkylferrocenes (alkyl = H, Me, Et, Bu and C5H11) by H2O2 to form the corresponding ferricenium cations catalyzed by horseradish peroxidase has been studied in micellar systems of Triton X-100, CTAB, and SDS, mostly at pH 6.0 and 25  °C. The rate of oxidation of ferrocenes with longer alkyl radicals is too slow to be measured. The reaction obeying the [RFc]:[H2O2] = 2 : 1 stoichiometry is strictly first-order in both HRP and RFc in a wide concentration range. The corresponding observed second-order rate constants k, which refer to the interaction of the peroxidase compound II (HRP-II) with RFc, decrease with the elongation of the alkyl substituent R, and this in turn is accompanied by an increase in the formal redox potentials E°′ in the same medium. Increasing the surfactant concentration lowers the rate constants k, the effect being due to the nonproductive binding of RFc to micelles rather than to enzyme inactivation. The micellar effects are accounted for in terms of the Berezin pseudo-phase model of micellar catalysis applied to the interaction of enzyme with organometallic substrates. The oxidation was found to occur primarily in the aqueous pseudo-phase and the calculated intrinsic second-order rate constants k w are (1.9 ± 0.5)×105, (2.7 ± 0.1)×104, and (5.9 ± 0.6)×103 M–1 s–1 for HFc, EtFc, and n–BuFc, respectively. The data obtained were used for estimating the self-exchange rate constants for the HRP-II/HRP couple in terms of the Marcus formalism. Received: 15 July 1996 / Accepted: 15 November 1996  相似文献   

10.
 Reactions (25  °C) of galactose oxidase, GOaseox from Fusarium NRRL 2903 with five different primary-alcohol-containing substrates RCH2OH:- D-galactose (I) and 2-deoxy-d-galactose (II) (monosaccharides); methyl-β-d-galactopyranoside (III) (glycoside);d-raffinose (IV) (trisaccharide); and dihydroxyacetone (V) have been studied in the presence of O2. The GOaseox state has a tyrosyl radical coordinated at a square-pyramidal CuII active site, and is a two-equivalent oxidant. Reactant concentrations were [GOaseox] (0.8–10 μM), RCH2OH (1.0–6.0 mM), and O2 (0.14–0.29 mM), with I=0.100 M (NaCl). The reactions, monitored at 450 nm by stopped-flow spectrophotometry, terminated with depletion of the O2. Each trace was fitted to the competing reactions GOaseox+RCH2 OH → GOaseredH2+RCHO (k 1), and GOaseredH2+O2→ GOaseox+H2O2 (k 2), with GOaseredH2 written as the doubly protonated two-electron-reduced CuI product. It was necessary to avoid auto-redox interconversion of GOaseox and GOasesemi . Information obtained at pH 7.5 indicates a 5 : 95 (ox : semi) "native" mix equilibration complete in ∼3 h. At pH >7.5, rate constants 10–4k 1 / M–1 s–1 for the reactions of GOaseox with (I) (1.19), (II) (1.07), (III) (1.29), (IV) (1.81), (V) (2.94) were determined. On decreasing the pH to 5.5, k 1 values decreased by factors of up to a half, and acid dissociation pK as in the range 6.6–6.9 were obtained. UV-Vis spectrophotometric studies on GOaseox gave an independently determined pK a of 6.7. No corresponding reactions of the Tyr495Phe variant were observed, and there are no similar UV-Vis absorbance changes for this variant. The pK a is therefore assigned to protonation of Tyr-495 which is a ligand to the Cu. The rate constant k 2 (1.01×107 M–1 s–1) is independent of pH in the range 5.5–9.0 investigated, suggesting that H+ (or H-atoms) for the O2 → H2O2 change are provided by the active site of GOasered . The CuI of GOasered is less extensively complexed, and a coordination number of three is likely. Received: 4 February 1997 / Accepted: 16 May 1997  相似文献   

11.
Copper(II) and nickel(II) complexes are prepared of potentially quadridentate ligands (LH3), N-{2-(2- hydroxyethylamino)ethyl}- (seeH3), N-{3-(2-hydroxy- ethylamino)propyl}- (steH3), and N-{2-(3-hydroxypropylamino)ethyl}-salicylamide (setH3). The nickel complexes Na[Ni(see)] and Na[Ni(set)]·1/2H2O are diamagnetic and square-planar, in which the ligands act as a quadridentate one coordinated through secondary amino-N, and deprotonated phenolic-O, alcoholic-O, and amido-N atoms. The three copper complexes Na [CuL]·H2O (L = see, ste, set) with a normal magnetic moment have a similar square-planar structure. In another type complexes Cu(LH)·H2O (LH = seeH, setH) an alcohol group is not deprotonated. Two isomers are present in Cu(seeH)·H2O: one has a normal and the other a subnormal magnetic moment. The difficulty of complex formation of steH3 may be attributed to an unfavourably fused 6-6-5 membered chelate ring with strain.  相似文献   

12.
 d(TpG) reacts with cis-[Pt(NH3)2(H2O)2]2+ in two steps to yield the platinum chelate cis-[Pt(NH3)2{d(TpG)-N3(1),N7(2)}]. In the latter, hindered rotation of the bases leads to an equilibrium between two rotamers interconverting slowly on the NMR time scale. The structure of the two rotameric chelates was studied by means of 1H NMR and molecular modeling techniques. The major and minor rotamers could be assigned unambiguously to the two head-to-head conformational domains which are characterized by syn/anti and anti/anti sugar-base orientations, respectively. Molecular models derived for both rotamers show that the orientations of the bases are mutually quasi-enantiomeric. The interconversion between the two rotamers (k ≈ 1 s–1 at 293 K) is approximately 104 times faster than the analogous rotamer interconversion observed in cis-[Pt(NH3)2{r(CpG)-N3(1),N7(2)}]+ [Girault J-P, Chottard G, Lallemand J-Y, Huguenin F, Chottard J-C (1984) J Am Chem Soc 106 : 7227–7232], suggesting that the steric clash of the exocyclic amino group of the platinum-bound cytosine with the ligands in cis position is more severe than that of the two thymine oxo groups. Received: 23 June 1997 / Accepted: 30 September 1997  相似文献   

13.
The copper(II), nickel(II) and zinc(II) binding ability of the multi-histidine peptide N-acetyl-His-Pro-His-His-NH2 has been studied by combined pH-potentiometry and visible, CD and EPR spectroscopies. The internal proline residue, preventing the metal ion induced successive amide deprotonations, resulted in the shift of this process toward higher pH values as compared to other peptides. The metal ions in the parent [ML]2+ complexes are exclusively bound by the three imidazole side chains. In [CuH−1L]+, formed between pH 6-8, the side chains of the two adjacent histidines and the peptide nitrogen between them are involved in metal ion binding. The next deprotonation results in the proton loss of the coordinated water molecule (CuH−1L(OH)). The latter two species exert polyfunctional catalytic activity, since they possess superoxide dismutase-, catecholase- (the oxidation of 3,5-di-tert-butylcatechol) and phosphatase-like (transesterification of the activated phosphoester 2-hydroxypropyl-4-nitrophenyl phosphate) properties. On further increase of the pH rearrangement of the coordination sphere takes place leading to the [CuH−3L] species, the deprotonated amide nitrogen displaces a coordinated imidazole nitrogen from the equatorial position of the metal ion. The shapes of the visible and CD spectra reflect a distorted arrangement of the donor atoms around the metal ion. In presence of zinc(II) the species [ZnL]2+ forms only above pH 6, which is shortly followed by precipitation. On the other hand, the [NiL]2+ complex is stable over a wide pH range, its deprotonation takes place only above pH 8. At pH 10 an octahedral NiH−2L species is present at first, which transforms slowly to a yellow square planar complex.  相似文献   

14.
 Fourier transform infrared (FTIR) spectroscopy is used to compare the thermally induced conformational changes in horse, bovine and tuna ferricytochromes c in 50 mM phosphate/0.2 M KCl. Thermal titration in D2O at pD 7.0 of the amide II intensity of the buried peptide NH protons reveals tertiary structural transitions at 54  °C in horse and at 57  °C in bovine c. These transitions, which occur well before loss of secondary structure, are associated with the alkaline isomerization involving Met80 heme-ligand exchange. In tuna c, the amide-II-monitored alkaline isomerization occurs at 35  °C, followed by a second amide II transition at 50  °C revealing a hitherto unreported conformational change in this cytochrome. Amide II transitions at 50  °C (tuna) and 54  °C (horse) are also observed during the thermal titration of the CN-ligated cytochromes (where CN displaces the Met80 ligand), but a well-defined 35  °C amide II transition is absent from the titration curve of the CNadduct of tuna c. The different mechanisms suggested by the FTIR data for the alkaline isomerization of tuna and the mammalian cytochromes c are discussed. After the alkaline isomerization, loss of secondary structure and protein aggregation occur within a 5  °C range with T m values at 74  °C (bovine c), 70  °C (horse c) and 65  °C (tuna c), as monitored by changes in the amide I′ bands. The FTIR spectra were also used to compare the secondary structures of the ferricytochromes c at 25  °C. Curve fitting of the amide I (H2O) and amide I′ (D2O) bands reveals essentially identical secondary structure in horse and bovine c, whereas splitting of the α-helical absorption of tuna c indicates the presence of less-stable helical structures. CN adduct formation results in no FTIR-detectable changes in the secondary structures of either tuna or horse c, indicating that Met80 ligation does not influence the secondary structural elements in these cytochromes. The data provided here demonstrate for the first time that the selective thermal titration of the amide II intensity of buried peptide NH protons in D2O is a powerful tool in protein conformational analysis. Received: 1 April 1999 / Accepted: 24 August 1999  相似文献   

15.
The 1:2 condensation of o-phenelenediamine and o-vanilline yields a compartmental N2O4 ligand N,N′-(1,2-Phenylene)-bis(3-methoxysalicylideneimine) [H2L]. When nickel(II) thiocyanate is added to the methanol solution of H2L, followed by addition of ammonium thiocyanate, an unusual nickel(II) compound, [NH4(NiL)2SCN]·H2O (1), is separated out in which an ammonium ion is sandwiched between two neutral square planner NiL moieties. Hydrogen bonding interactions are observed among the ammonium ion, NiL moieties, the thiocyanate anion and the water of crystallization. The compound is characterized by C, H, N analysis, UV-VIS and IR spectroscopy, room temperature magnetic susceptibility measurement and X-ray crystal diffraction study. The compound crystallizes in monoclinic space group P21/n with a = 13.8636(7) Å, b = 14.0267(7) Å, c = 22.2715(10) Å and β = 94.301(3)°.  相似文献   

16.
The solid-state structures of 6-(1-methylbenzimidazol-2-yl)-1H-pyridin-2-one (HL) and the copper(II) complex [Cu(L)2(OH2)]·2H2O (1) are established by X-ray crystallography and also by means of physicochemical and spectroscopic methods. The molecules of HL form a self-complementary head-to-tail hydrogen-bonded dimer through C-H?N and C-H?O contacts to give an infinite 1D chain. The copper(II) complex (1) is five-coordinate with distorted trigonal-bipyramidal (TBP) geometry of the N4O donor atoms. The electronic and EPR data are in agreement with the X-ray structure of 1, showing that HL coordinates to copper(II) centre as a mono-anionic ligand through deprotonated pyridone N atom and the tertiary benzimidazole nitrogen atom to form a neutral complex in which the water molecule occupies the fifth position. The 1D water tape, T4(2)7(2)6(2)7(2) is anchored to the host through hydrogen bonds between coordinated water molecule [O(3w)] as acting double H-donor, pyridone carbonyl groups [O(2) and O(1)] as double H-acceptor and the lattice water molecules [O(4w) and O(5w)] as double H-donor and single H-acceptor).  相似文献   

17.
 Four reductions of the R2 subunit of mouse ribonucleotide reductase have been studied and found to exhibit different behaviour from that of Escherichia coli R2. An important difference is that there is no stable met-R2 (Fe2 II I) form of mouse R2. With hydroxyurea, hydrazine and hydroxylamine uniphasic kinetics are observed for the combined reduction of radical Tyr ˙ and Fe2 II I components to tyrosine and Fe2 II respectively. The rate constants, determined at 370 nm (emphasising FeIII decay) and 417 nm (emphasising Tyr ˙ decay), differ by factors of 2–3, allowing some mechanistic features to be defined. The studies with hydrazine are particularly important. In the case of E. coli R2, a first phase corresponding to two-equivalent reduction of the met-R2 component has been observed [18]. It is likely that the four times slower second phase reaction of active E. coli R2 also corresponds to the Fe2 II I → Fe2 II change and is followed by fast intramolecular Fe2 II reduction of the higher potential Tyr ˙. The latter changes are believed to hold also for (active) mouse R2. The FeIIFeIII semi forms have been detected at low levels by EPR for mouse R2 (9%) and E. coli (∼5%) in previous studies. Further substrate reduction of FeIIFeIII occurs at a comparable rate to account for the transient behaviour of FeIIFeIII. For mouse R2 the combined FeIII decay processes (which we are unable to separate) give smaller uniphasic rate constants at 370 nm than at 417 nm. A fitted-base-line (FBL) treatment of absorbance changes at 417 nm targets more closely the Tyr ˙ decay as a means of monitoring the rate-determining step. The FBL method gives rate constants k (M–1 s–1) at 25  °C and pH 7.5 for hydroxyurea (1.46), hydrazine (0.163) and hydroxylamine (4.4). Surprisingly, phenylhydrazine, with a less strong reduction potential (0.25 V), gives a substantially faster reduction of the Tyr ˙ as the only redox step (rate constant 27 M–1 s–1). In this case a slower second phase at 370 nm is independent of reductant and corresponds to rate-controlling release of FeIII. Overall the results indicate a more reactive redox centre for mouse R2 and help develop further an understanding of factors affecting the reactivity of R2. Received: 11 October 1996 / Accepted: 11 February 1997  相似文献   

18.
Combined pH-metric, UV-Vis, 1H NMR and EPR spectral investigations on the complex formation of M(II) ions (M=Co, Ni, Cu and Zn) with N-(2-benzimidazolyl)methyliminodiacetic acid (H2bzimida, hereafter H2L) in aqueous solution at a fixed ionic strength, I=10−1 mol dm−3, at 25 ± 1 °C indicate the formation of M(L), M(H−1L) and M2(H−1L)+ complexes. Proton-ligand and metal-ligand constants and the complex formation equilibria have been elucidated. Solid complexes, [M(L)(H2O)2] · nH2O (n=1 for M = Co and Zn, n=2 for M = Ni) and {Cu (μ-L) · 4H2O}n, have been isolated and characterized by elemental analysis, spectral, conductance and magnetic measurements and thermal studies. Structures of [Ni(L)(H2O)2] · 2H2O and {Cu(μ-L) · 4H2O}n have been determined by single crystal X-ray diffraction. The nickel(II) complex exists in a distorted octahedral environment in which the metal ion is coordinated by the two carboxylate O atoms, the amino-N atom of the iminodiacetate moiety and the pyridine type N-atom of the benzimidazole moiety. Two aqua O atoms function as fifth and sixth donor atoms. The copper(II) complex is made up of interpenetrating polymeric chains of antiferromagnetically coupled Cu(II) ions linked by carboxylato bridges in syn-anti (apical-equatorial) bonding mode and stabilized via interchain hydrogen bonds and π-π stacking interactions.  相似文献   

19.
Despite the availability of many mutants for signal transduction, Arabidopsis thaliana guard cells have so far not been used in electrophysiological research. Problems with the isolation of epidermal strips and the small size of A. thaliana guard cells were often prohibiting. In the present study these difficulties were overcome and guard cells were impaled with double-barreled microelectrodes. Membrane-potential recordings were often stable for over half an hour and voltage-clamp measurements could be conducted. The guard cells were found to exhibit two states. The majority of the guard cells had depolarized membrane potentials, which were largely dependent on external K+ concentrations. Other cells displayed spontaneous transitions to a more hyperpolarized state, at which the free-running membrane potential (Em) was not sensitive to the external K+ concentration. Two outward-rectifying conductances were identified in cells in the depolarized state. A slow outward-rectifying channel (s-ORC) had properties resembling the K+-selective ORC of Vicia faba guard cells (Blatt, 1988, J Membr Biol 102: 235–246). The activation and inactivation times and the activation potential, all depended on the reversal potential (Erev) of the s-ORC conductance. The s-ORC was blocked by Ba2+ (K1/2 = 0.3–1.3mM) and verapamil (K1/2 = 15–20 μM). A second rapid outward-rectifying conductance (r-ORC) activated instantaneously upon stepping the voltage to positive values and was stimulated by Ba2+. Inward-rectifying channels (IRC) were only observed in cells in the hyperpolarized state. The activation time and activation potential of this channel were not sensitive to the external K+ concentration. The slow activation of the IRC (t1/2 ≈ 0.5 s) and its negative activation potential (Vthreshold = −155 mV) resemble the values found for the KAT1 channel expressed in Saccharomyces cerevisiae (Bertl et al., 1995, Proc Natl Acad Sci USA 92: 2701–2705). The results indicate that A. thaliana guard cells provide an excellent system for the study of signal transduction processes. Received: 28 March 1996 / Accepted: 11 November 1996  相似文献   

20.
The freshwater microalga Haematococcus pluvialis is one of the best microbial sources of the carotenoid astaxanthin, but this microalga shows low growth rates and low final cell densities when cultured with traditional media. A single-variable optimization strategy was applied to 18 components of the culture media in order to maximize the productivity of vegetative cells of H. pluvialis in semicontinuous culture. The steady-state cell density obtained with the optimized culture medium at a daily volume exchange of 20% was 3.77 · 105 cells ml−1, three times higher than the cell density obtained with Bold basal medium and with the initial formulation. The formulation of the optimal Haematococcus medium (OHM) is (in g l−1) KNO3 0.41, Na2HPO4 0.03, MgSO4 · 7H2O 0.246, CaCl2 · 2H2O 0.11, (in mg l−1) Fe(III)citrate · H2O 2.62, CoCl2 · 6H2O 0.011, CuSO4 · 5H2O 0.012, Cr2O3 0.075, MnCl2 · 4H2O 0.98, Na2MoO4 · 2H2O 0.12, SeO2 0.005 and (in μg l−1]) biotin 25, thiamine 17.5 and B12 15. Vanadium, iodine, boron and zinc were demonstrated to be non-essential for the growth of H. pluvialis. Higher steady-state cell densities were obtained by a three-fold increase of all nutrient concentrations but a high nitrate concentration remained in the culture medium under such conditions. The high cell productivities obtained with the new optimized medium can serve as a basis for the development of a two-stage technology for the production of astaxanthin from H. pluvialis. Received: 10 September 1999 / Received revision: 2 December 1999 / Accepted: 3 December 1999  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