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 goal of our work was a throughout characterization of the pharmacology of the TIPP-analog, Tyr-Tic-(2S,3R)-β-MePhe-Phe-OH and see if putative δ-opioid receptor subtypes can be distinguished. Analgesic latencies were assessed in mouse tail-flick assays after intrathecal administration. In vitro receptor autoradiography, binding and ligand-stimulated [35S]GTPγS functional assays were performed in the presence of putative δ1-(DPDPE: agonist, BNTX: antagonist), δ2-(agonist: deltorphin II, Ile5,6-deltorphin II, antagonist: naltriben) and μ-(DAMGO: agonist) opioid ligands. The examined antagonist inhibited the effect of DPDPE by 60%, but did not antagonize δ2- and μ-agonist induced analgesia. The radiolabeled form identified binding sites with KD = 0.18 nM and receptor densities of 102.7 fmol/mg protein in mouse brain membranes. The binding site distribution of the [3H]Tyr-Tic-(2S,3R)-β-MePhe-Phe-OH agreed well with that of [3H]Ile5,6-deltorphin II as revealed by receptor autoradiography. Tyr-Tic-(2S,3R)-β-MePhe-Phe-OH displayed 2.49 ± 0.06 and 0.30 ± 0.01 nM potency against DPDPE and deltorphin II in the [35S]GTPγS functional assay, respectively. The rank order of potency of putative δ1- and δ2-antagonists against DPDPE and deltorphin was similar in brain and CHO cells expressing human δ-opioid receptors. Deletion of the DOR-1 gene resulted in no residual binding of the radioligand and no significant DPDPE effect on G-protein activation. Tyr-Tic-(2S,3R)-β-MePhe-Phe-OH is a highly potent and δ-opioid specific antagonist both in vivo and in vitro. However, the putative δ1- and δ2-opioid receptors could not be unequivocally distinguished in vitro.  相似文献   

2.
A case study on Centaurea gymnocarpa Moris & De Not., a narrow endemic species, was carried out by analyzing its morphological, anatomical, and physiological traits in response to natural habitat stress factors under Mediterranean climate conditions. The results underline that the species is particularly adapted to the environment where it naturally grows. At the plant level, the above-ground/below-ground dry mass (1.73 ± 0.60) shows its investment predominately in the above-ground structure with a resulting total leaf area per plant of 1399 ± 94 cm2. The senescent attached leaves at the base of the plant contribute to limit leaf transpiration by shading soil around the plant. Moreover, the dense C. gymnocarpa leaf pubescence, leaf rolling, the relatively high leaf mass area (LMA = 12.3 ± 1.3 mg cm−2) and leaf tissue density (LTD = 427 ± 44 mg cm−3) contribute to limit leaf transpiration, also postponing leaf death under dry conditions. At the physiological level, a relatively low respiration/photosynthesis ratio (R/PN) in spring results from high R [2.26 ± 0.59 μmol (CO2) m−2 s−1] and PN [12.3 ± 1.5 μmol (CO2) m−2 s−1]. The high photosynthetic nitrogen use efficiency [PNUE = 15.5 ± 0.4 μmol (CO2) g−1 (N) s−1] shows the large amount of nitrogen (N) invested in the photosynthetic machinery of new leaves, associated to a high chlorophyll content (Chl = 35 ± 5 SPAD units). On the contrary, the highest R/PN ratio (1.75 ± 0.19) in summer is due to a significant PN decrease and increase of R in response to drought. The low PNUE [1.5 ± 0.2 μmol (CO2) g−1 (N) s−1] in this season is indicative of a greater N investment in leaf cell walls which may contribute to limit transpiration. On the contrary, the low R/PN ratio (0.05 ± 0.02) in winter is resulting from the limited enzyme activity of the respiratory apparatus [R = 0.23 ± 0.08 μmol (CO2) m−2 s−1] while the low PNUE [3.5 ± 0.2 μmol (CO2) g−1 (N) s−1] suggests that low temperatures additionally limit plant production. The experiment of the imposed water stress confirms that the C. gymnocarpa growth capability is in conformity with the severe conditions of its natural habitat, likewise as it may be the case with others narrow endemic species that have occupied niches with similar extreme conditions.  相似文献   

3.
We examine hemolymph ion regulation and the kinetic properties of a gill microsomal (Na+, K+)-ATPase from the intertidal hermit crab, Clibanarius vittatus, acclimated to 45‰ salinity for 10 days. Hemolymph osmolality is hypo-regulated (1102.5 ± 22.1 mOsm kg−1 H2O) at 45‰ but elevated compared to fresh-caught crabs (801.0 ± 40.1 mOsm kg−1 H2O). Hemolymph [Na+] (323.0 ± 2.5 mmol L−1) and [Mg2+] (34.6 ± 1.0 mmol L−1) are hypo-regulated while [Ca2+] (22.5 ± 0.7 mmol L−1) is hyper-regulated; [K+] is hyper-regulated in fresh-caught crabs (17.4 ± 0.5 mmol L−1) but hypo-regulated (6.2 ± 0.7 mmol L−1) at 45‰. Protein expression patterns are altered in the 45‰-acclimated crabs, although Western blot analyses reveal just a single immunoreactive band, suggesting a single (Na+, K+)-ATPase α-subunit isoform, distributed in different density membrane fractions. A high-affinity (Vm = 46.5 ± 3.5 U mg−1; K0.5 = 7.07 ± 0.01 μmol L−1) and a low-affinity ATP binding site (Vm = 108.1 ± 2.5 U mg−1; K0.5 = 0.11 ± 0.3 mmol L−1), both obeying cooperative kinetics, were disclosed. Modulation of (Na+, K+)-ATPase activity by Mg2+, K+ and NH4+ also exhibits site-site interactions, but modulation by Na+ shows Michaelis-Menten kinetics. (Na+, K+)-ATPase activity is synergistically stimulated up to 45% by NH4+ plus K+. Enzyme catalytic efficiency for variable [K+] and fixed [NH4+] is 10-fold greater than for variable [NH4+] and fixed [K+]. Ouabain inhibited ≈80% of total ATPase activity (KI = 464.7 ± 23.2 μmol L−1), suggesting that ATPases other than (Na+, K+)-ATPase are present. While (Na+, K+)-ATPase activities are similar in fresh-caught (around 142 nmol Pi min−1 mg−1) and 45‰-acclimated crabs (around 154 nmol Pi min−1 mg−1), ATP affinity decreases 110-fold and Na+ and K+ affinities increase 2-3-fold in 45‰-acclimated crabs.  相似文献   

4.
In this paper we explore the use of fluorescently labeled cytochrome c peroxidase (CcP) from baker's yeast for monitoring nitric oxide (NO) down to the sub-micromolar level, by means of a FRET (Förster Resonance Energy Transfer) mechanism. The binding affinity constant (Kd) for the NO binding to CcP was determined to be 10 ± 1.5 µM. The rate of NO dissociation from the CcP (koff) and the second order rate constant for the NO association (kon) were found to be 0.22 ± 0.08 min− 1 and 0.024 ± 0.002 µM− 1 min− 1 respectively. The immobilization of fluorescently labeled CcP into a polymeric matrix for use in a solid state NO sensing device was also explored. The results provide proof-of-principle that labeled CcP can be successfully implemented in a fast, simple, quantitative and sensitive NO sensing device.  相似文献   

5.
The transport of nickel (Ni) across the renal brush border membrane of the rainbow trout was examined in vitro using brush border membrane vesicles (BBMVs). Both transmembrane transport of Ni into an osmotically active intravesicular space, and binding of Ni to the brush border membrane itself, were confirmed. Nickel (Ni) uptake fitted a two component kinetic model. Saturable, temperature-dependent transport dominated at lower Ni concentrations, with a moderate linear diffusive component of Ni transport apparent at higher Ni concentrations. An affinity constant (Km) for Ni transport within the specifically described vesicular media was calculated as 17.9 ± 1.9 μM, the maximal rate of transport (Jmax) was calculated as 108.3 ± 3.7 nmol mg protein−1 min−1, and the slope of the linear diffusive component was calculated as 0.049 ± 0.005 nmol mg protein−1 min−1 per μM of Ni. Efflux of Ni from BBMVs was fitted to an exponential decay curve with a half-time (T1/2) of 125.2 ± 7.3 s. Ni uptake into renal BBMVs was inhibited by magnesium at a 100:1 Mg to Ni molar ratio, and by magnesium and calcium at a 1000:1 molar ratio. In the presence of histidine at a 100:1 histidine to Ni ratio, Ni uptake was almost completely abolished. At a 1:1 molar ratio, histidine inhibited Ni uptake by approximately 50%. Ni-histidine complexation was rapid, with a T1/2 of 12.2 s describing the Ni-histidine equilibration time needed to inhibit Ni uptake into renal BBMVs by 50%. Characterization of Ni transport across cellular membranes is an important step in understanding both the processes underlying homeostatic regulation of Ni, and the toxicological implications of excessive Ni exposure in aquatic ecosystems.  相似文献   

6.
The interaction of (−)-reboxetine, a non-tricyclic norepinephrine selective reuptake inhibitor, with muscle-type nicotinic acetylcholine receptors (AChRs) in different conformational states was studied by functional and structural approaches. The results established that (−)-reboxetine: (a) inhibits (±)-epibatidine-induced Ca2+ influx in human (h) muscle embryonic (hα1β1γδ) and adult (hα1β1εδ) AChRs in a non-competitive manner and with potencies IC50 = 3.86 ± 0.49 and 1.92 ± 0.48 μM, respectively, (b) binds to the [3H]TCP site with ∼13-fold higher affinity when the Torpedo AChR is in the desensitized state compared to the resting state, (c) enhances [3H]cytisine binding to the resting but activatableTorpedo AChR but not to the desensitized AChR, suggesting desensitizing properties, (d) overlaps the PCP luminal site located between rings 6′ and 13′ in the Torpedo but not human muscle AChRs. In silico mutation results indicate that ring 9′ is the minimum structural component for (−)-reboxetine binding, and (e) interacts to non-luminal sites located within the transmembrane segments from the Torpedo AChR γ subunit, and at the α1/ε transmembrane interface from the adult muscle AChR. In conclusion, (−)-reboxetine non-competitively inhibits muscle AChRs by binding to the TCP luminal site and by inducing receptor desensitization (maybe by interacting with non-luminal sites), a mechanism that is shared by tricyclic antidepressants.  相似文献   

7.
A combination of isothermal titration calorimetry (ITC), topoisomerase I DNA unwinding assays, and ethidium bromide displacement studies were employed to investigate the binding of a homologous series of naphthalene diimides (NDI) to DNA. Our results suggest that the nature of the substituent plays a significant role in both the preferred binding mode and relative binding affinity of the compounds of this study. Only intercalative-type binding (K = 15 ± 3 × 106 M−1) was observed for the NDI with the smallest substituent (trimethyl-ethylamino), while larger members of the series (diethylmethyl-, dipropylmethyl- and dibutylmethyl-ethylamino substituents) adopted an additional binding mode of higher affinity (K1 = 31 − 78 × 106 M−1).  相似文献   

8.
Reaction of the five-coordinate trigonal-bipyramidal platinum(II) complex, [Pt(pt)(pp3)](BF4) (pt = 1-propanethiolate, pp3 = tris[2-(diphenylphosphino)ethyl]phosphine), with I in chloroform gave the five-coordinate square-pyramidal complex with a dissociated terminal phosphino group and an apically coordinated iodide ion in equilibrium. The thermodynamic parameters for the equilibrium between the trigonal-bipyramidal and square-pyramidal geometries, [Pt(pt)(pp3)]+ + I ? [PtI(pt) (pp3)], and the kinetic parameters for the chemical exchange were obtained as follows: , ΔH0 = − 10 ± 2.4 kJ mol−1, ΔS0 = − 36 ± 10 J K−1 mol−1, , ΔH = 34 ± 4.7 kJ mol−1, ΔS = − 50 ± 21 J K−1 mol−1. The square-planar trinuclear platinum(II) complex was formed by bridging reaction of one of the terminal phosphino groups of trigonal-bipyramidal [PtCl(pp3)]Cl with trans-[PtCl2(NCC6H5)2] in chloroform. From these facts, ligand substitution reactions of [PtX(pp3)]+ (X = monodentate anion) are expected to proceed via an intermediate with a dissociated phosphino group. The rate constants for the chloro-ligand substitution reactions of [PtCl(pp3)]+ with Br and I in chloroform approached the respective limiting values as concentrations of the entering halide ions are increased. These kinetic results confirmed the preassociation mechanism in which the square pyramidal intermediate with a dissociated phosphino group and an apically coordinated halide ion is present in the rapid pre-equilibrium.  相似文献   

9.
To characterize driving forces and driven processes in formation of a large-interface, wrapped protein-DNA complex analogous to the nucleosome, we have investigated the thermodynamics of binding the 34-base pair (bp) H′ DNA sequence to the Escherichia coli DNA-remodeling protein integration host factor (IHF). Isothermal titration calorimetry and fluorescence resonance energy transfer are applied to determine effects of salt concentration [KCl, KF, K glutamate (KGlu)] and of the excluded solute glycine betaine (GB) on the binding thermodynamics at 20 °C. Both the binding constant Kobs and enthalpy ΔH°obs depend strongly on [salt] and anion identity. Formation of the wrapped complex is enthalpy driven, especially at low [salt] (e.g., ΔHoobs = − 20.2 kcal·mol− 1 in 0.04 M KCl). ΔH°obs increases linearly with [salt] with a slope (dΔH°obs/d[salt]), which is much larger in KCl (38 ± 3 kcal·mol− 1 M− 1) than in KF or KGlu (11 ± 2 kcal·mol− 1 M− 1). At 0.33 M [salt], Kobs is approximately 30-fold larger in KGlu or KF than in KCl, and the [salt] derivative SKobs = dlnKobs/dln[salt] is almost twice as large in magnitude in KCl (− 8.8 ± 0.7) as in KF or KGlu (− 4.7 ± 0.6).A novel analysis of the large effects of anion identity on Kobs, SKobs and on ΔH°obs dissects coulombic, Hofmeister, and osmotic contributions to these quantities. This analysis attributes anion-specific differences in Kobs, SKobs, and ΔH°obs to (i) displacement of a large number of water molecules of hydration [estimated to be 1.0(± 0.2) × 103] from the 5340 Å2 of IHF and H′ DNA surface buried in complex formation, and (ii) significant local exclusion of F and Glu from this hydration water, relative to the situation with Cl, which we propose is randomly distributed. To quantify net water release from anionic surface (22% of the surface buried in complexation, mostly from DNA phosphates), we determined the stabilizing effect of GB on Kobs: dlnKobs/d[GB]  = 2.7 ± 0.4 at constant KCl activity, indicating the net release of ca. 150 H2O molecules from anionic surface.  相似文献   

10.
Kinetic studies of X exchange on [AuX4] square-planar complexes (where X=Cl and CN) were performed at acidic pH in the case of chloride system and as a function of pH for the cyanide one. Chloride NMR study (330-365 K) gives a second-order rate law on [AuCl4] with the kinetic parameters: (k2Au,Cl)298=0.56±0.03 s−1 mol−1 kg; ΔH2‡ Au,Cl=65.1±1 kJ mol−1; ΔS2‡ Au,Cl=−31.3±3 J mol−1 K−1 and ΔV2 Au,Cl=−14±2 cm3 mol−1. The variable pressure data clearly indicate the operation of an Ia or A mechanism for this exchange pathway. The proton exchange on HCN was determined by 13C NMR as a function of pH and the rate constant of the three reaction pathways involving H2O, OH and CN were determined: k0HCN,H=113±17 s−1, k1HCN,H=(2.9±0.7)×109 s−1 mol−1 kg and k2HCN,H=(0.6±0.2)×106 s−1 mol−1 kg at 298.1 K. The rate law of the cyanide exchange on [Au(CN)4] was found to be second order with the following kinetic parameters: (k2Au,CN)298=6240±85 s−1 mol−1 kg, ΔH2 Au,CN=40.0±0.8 kJ mol−1, ΔS2 Au,CN=−37.8±3 J mol−1 K−1 and ΔV2 Au,CN=+2±1 cm3 mol−1. The rate constant observed varies about nine orders of magnitude depending on the pH and HCN does not act as a nucleophile. The observed rate constant of X exchange on [AuX4] are two or three orders of magnitude faster than the Pt(II) analogue.  相似文献   

11.
Temperature-sensitive liposomes (TSLs) loaded with doxorubicin (Dox), and Magnetic Resonance Imaging contrast agents (CAs), either manganese (Mn2 +) or [Gd(HPDO3A)(H2O)], provide the advantage of drug delivery under MR image guidance. Encapsulated MRI CAs have low longitudinal relaxivity (r1) due to limited transmembrane water exchange. Upon triggered release at hyperthermic temperature, the r1 will increase and hence, provides a means to monitor drug distribution in situ. Here, the effects of encapsulated CAs on the phospholipid bilayer and the resulting change in r1 were investigated using MR titration studies and 1H Nuclear Magnetic Relaxation Dispersion (NMRD) profiles. Our results show that Mn2 + interacted with the phospholipid bilayer of TSLs and consequently, reduced doxorubicin retention capability at 37 °C within the interior of the liposomes over time. Despite that, Mn2 +-phospholipid interaction resulted in higher r1 increase, from 5.1 ± 1.3 mM− 1 s− 1 before heating to 32.2 ± 3 mM− 1 s− 1 after heating at 60 MHz and 37 °C as compared to TSL(Gd,Dox) where the longitudinal relaxivities before and after heating were 1.2 ± 0.3 mM− 1 s− 1 and 4.4 ± 0.3 mM− 1 s− 1, respectively. Upon heating, Dox was released from TSL(Mn,Dox) and complexation of Mn2 + to Dox resulted in a similar Mn2 + release profile. From 25 to 38 °C, r1 of [Gd(HPDO3A)(H2O)] gradually increased due to increase transmembrane water exchange, while no Dox release was observed. From 38 °C, the release of [Gd(HPDO3A)(H2O)] and Dox was irreversible and the release profiles coincided. By understanding the non-covalent interactions between the MRI CAs and phospholipid bilayer, the properties of the paramagnetic TSLs can be tailored for MR guided drug delivery.  相似文献   

12.
Two 15N-labelled cis-Pt(II) diamine complexes with dimethylamine (15N-dma) and isopropylamine (15N-ipa) ligands have been prepared and characterised. [1H,15N] HSQC NMR spectroscopy is used to obtain the rate and equilibrium constants for the aquation of cis-[PtCl2(15N-dma)2] at 298 K in 0.1 M NaClO4 and to determine the pKa values of cis-[PtCl(H2O)(15N-dma)2]+ (6.37) and cis-[Pt(H2O)2(15N-dma)2]2+ (pKa1 = 5.17, pKa2 = 6.47). The rate constants for the first and second aquation steps (k1 = (2.12 ± 0.01) × 10−5 s−1, k2 = (8.7 ± 0.7) × 10−6 s−1) and anation steps (k−1 = (6.7 ± 0.8) × 10−3 M−1 s−1, k−2 = 0.043 ± 0.004 M−1 s−1) are very similar to those reported for cisplatin under similar conditions, and a minor difference is that slow formation of the hydroxo-bridged dimer is observed. Aquation studies of cis-[PtCl2(15N-ipa)2] were precluded by the close proximity of the NH proton signal to the 1H2O resonance.  相似文献   

13.
The Iberian Peninsula encompasses more than 80% of the species richness of European aquatic ranunculi. The floristic diversity of the phytocoenosis characterised by aquatic Ranunculus and the main physical–chemical factors of the water were studied in 43 localities of the central Iberian Peninsula. Four aquatic Ranunculus communities are found in most of the aquatic environments. These are species-poor and have an uneven distribution: three species of Batrachium are heterophyllous and their communities are distributed in different aquatic ecosystems on silicated substrates; one species is homophyllous and its community occurs in various aquatic ecosystems with carbonated waters. In the Mediterranean climate, Ranunculus species are present in different habitats, as shown by the results of all the statistical analyses. Ranunculus trichophyllus communities occur in base-rich waters with a high buffering capacity (2273.44 ± 794.57 mg CaCO3 L−1) and a high concentration of cations (Ca2+, 121 ± 33.12 mg L−1; Mg2+, 71.64 ± 82.77 mg L−1), nitrates (2.89 ± 4.80 mg L−1), ammonium (2.19 ± 1.36 mg L−1) and sulphates (216.25 ± 218.54 mg L−1). Ranunculus penicillatus communities are found in flowing waters with a high concentration of phosphates (0.48 ± 0.6 mg L−1) and intermediate buffering capacity (683.66 ± 446.76 mg CaCO3 L−1). Both Ranunculus pseudofluitans and Ranunculus peltatus communities grow in waters with low buffering capacity (R. pseudofluitans, 385.91 ± 209.2 mg CaCO3 L−1; R. peltatus, 263.3 ± 180.36 mg CaCO3 L−1), and a low concentration of cations (R. pseudofluitans: Ca2+, 12.57 ± 9.42 mg L−1; Mg2+, 3.42 ± 1.67 mg L−1; R. peltatus: Ca2+, 15 ± 18.26 mg L−1; Mg2+, 6.26 ± 8.89 mg L−1) and nutrients (R. pseudofluitans: nitrates, 0.23 ± 0.2 mg L−1; phosphates, 0.09 ± 0.1 mg L−1; R. peltatus: nitrates, 0.19 ± 0.21 mg L−1; phosphates, 0.09 ± 0.12 mg L−1); the first in flowing waters, the latter in still waters.  相似文献   

14.
The binding affinity of the two substrate–water molecules to the water-oxidizing Mn4CaO5 catalyst in photosystem II core complexes of the extremophilic red alga Cyanidioschyzon merolae was studied in the S2 and S3 states by the exchange of bound 16O-substrate against 18O-labeled water. The rate of this exchange was detected via the membrane-inlet mass spectrometric analysis of flash-induced oxygen evolution. For both redox states a fast and slow phase of water-exchange was resolved at the mixed labeled m/z 34 mass peak: kf = 52 ± 8 s− 1 and ks = 1.9 ± 0.3 s− 1 in the S2 state, and kf = 42 ± 2 s− 1 and kslow = 1.2 ± 0.3 s− 1 in S3, respectively. Overall these exchange rates are similar to those observed previously with preparations of other organisms. The most remarkable finding is a significantly slower exchange at the fast substrate–water site in the S2 state, which confirms beyond doubt that both substrate–water molecules are already bound in the S2 state. This leads to a very small change of the affinity for both the fast and the slowly exchanging substrates during the S2 → S3 transition. Implications for recent models for water-oxidation are briefly discussed.  相似文献   

15.
Sulfonylurea drugs are often prescribed as a treatment for type II diabetes to help lower blood sugar levels by stimulating insulin secretion. These drugs are believed to primarily bind in blood to human serum albumin (HSA). This study used high-performance affinity chromatography (HPAC) to examine the binding of sulfonylureas to HSA. Frontal analysis with an immobilized HSA column was used to determine the association equilibrium constants (Ka) and number of binding sites on HSA for the sulfonylurea drugs acetohexamide and tolbutamide. The results from frontal analysis indicated HSA had a group of relatively high-affinity binding regions and weaker binding sites for each drug, with average Ka values of 1.3 (±0.2) × 105 and 3.5 (±3.0) × 102 M−1 for acetohexamide and values of 8.7 (±0.6) × 104 and 8.1 (±1.7) × 103 M−1 for tolbutamide. Zonal elution and competition studies with site-specific probes were used to further examine the relatively high-affinity interactions of these drugs by looking directly at the interactions that were occurring at Sudlow sites I and II of HSA (i.e., the major drug-binding sites on this protein). It was found that acetohexamide was able to bind at both Sudlow sites I and II, with Ka values of 1.3 (±0.1) × 105 and 4.3 (±0.3) × 104 M−1, respectively, at 37 °C. Tolbutamide also appeared to interact with both Sudlow sites I and II, with Ka values of 5.5 (±0.2) × 104 and 5.3 (±0.2) × 104 M−1, respectively. The results provide a more quantitative picture of how these drugs bind with HSA and illustrate how HPAC and related tools can be used to examine relatively complex drug–protein interactions.  相似文献   

16.
The ruthenium(II) hexaaqua complex [Ru(H2O)6]2+ reacts with dihydrogen under pressure to give the η2-dihydrogen ruthenium(II) pentaaqua complex [Ru(H2)(H2O)5]2+.The complex was characterized by 1H, 2H and 17O NMR: δH = −7.65 ppm, JHD = 31.2 Hz, δO = −80.4 ppm (trans to H2) and δO = −177.4 ppm (cis to H2).The H-H distance in coordinated dihydrogen was estimated to 0.889 Å from JHD, which is close to the value obtained from DFT calculations (0.940 Å).Kinetic studies were performed by 1H and 2H NMR as well as by UV-Vis spectroscopy, yielding the complex formation rate and equilibrium constants: kf = (1.7 ± 0.2) × 10−3 kg mol−1 s−1 and Keq = 4.0 ± 0.5 mol kg−1.The complex formation rate with dihydrogen is close to values reported for other ligands and thus it is assumed that the reaction with dihydrogen follows the same mechanisn (Id).In deuterated water, one can observe that [Ru(H2)(H2O)5]2+ catalyses the hydrogen exchange between the solvent and the dissolved dihydrogen.A hydride is proposed as the intermediate for this exchange.Using isotope labeling, the rate constant for the hydrogen exchange on the η2-dihydrogen ligand was determined as k1 = (0.24 ± 0.04) × 10−3 s−1.The upper and lower limits of the pKa of the coordinated dihydrogen ligand have been estimated:3 < pKa < 14.  相似文献   

17.
Human serum albumin (HSA) is a monomeric allosteric protein. Here, the effect of ibuprofen on denitrosylation kinetics (koff) and spectroscopic properties of HSA-heme-Fe(II)-NO is reported. The koff value increases from (1.4 ± 0.2) × 10−4 s−1, in the absence of the drug, to (9.5 ± 1.2) × 10−3 s−1, in the presence of 1.0 × 10−2 M ibuprofen, at pH 7.0 and 10.0 °C. From the dependence of koff on the drug concentration, values of the dissociation equilibrium constants for ibuprofen binding to HSA-heme-Fe(II)-NO (K1 = (3.1 ± 0.4) × 10−7 M, K2 = (1.7 ± 0.2) × 10−4 M, and K3 = (2.2 ± 0.2) × 10−3 M) were determined. The K3 value corresponds to the value of the dissociation equilibrium constant for ibuprofen binding to HSA-heme-Fe(II)-NO determined by monitoring drug-dependent absorbance spectroscopic changes (H = (2.6 ± 0.3) × 10−3 M). Present data indicate that ibuprofen binds to the FA3-FA4 cleft (Sudlow’s site II), to the FA6 site, and possibly to the FA2 pocket, inducing the hexa-coordination of HSA-heme-Fe(II)-NO and triggering the heme-ligand dissociation kinetics.  相似文献   

18.
Homoleptic eight- and nine-coordinate U(IV) perchlorate complexes with sulfoxide ligands have been characterized crystallographically. Crystals of [U(dmso)8](ClO4)4 · 0.75CH3NO2, [U(dmso)9](ClO4)4 · 4dmso (dmso = dimethyl sulfoxide), and [U(tmso)8](ClO4)4 · 2tmso (tmso = tetramethylene sulfoxide) were found to have dodecahedral, tricapped trigonal prismatic, and square antiprismatic geometries, respectively. Average U-O bond distances in [U(dmso)8](ClO4)4 · 0.75CH3NO2, [U(dmso)9](ClO4)4 · 4dmso, and [U(tmso)8](ClO4)4 · 2tmso are 2.35(3), 2.41 (4), and 2.35(3) Å, respectively. Furthermore, it was found that [U(dmso)8]4+ is in equilibrium with [U(dmso)9]4+ in CH3NO2 solution containing dmso. Thermodynamic parameters for such an equilibrium are as follows: K (25 °C) = 3.4 ± 0.2 dm3 mol−1, ΔH = −54.9 ± 4.5 kJ mol−1, and ΔS = −174 ± 15 J K−1 mol−1.  相似文献   

19.
Low concentrations of urea and GuHCl (2 M) enhanced the activity of endoglucanase (EC 3.1.2.4) from Aspergillus aculeatus by 2.3- and 1.9-fold, respectively. The Km values for controls, in the presence of 2 M urea and GuHCl, were found to be 2.4 ± 0.2 × 10−8 mol L−1, 1.4 ± 0.2 × 10−8 mol L−1, and 1.6 ± 0.2 × 10−8 mol L−1, respectively. The dissociation constant (Kd) showed changes in the affinity of the enzyme for the substrate with increases in the Kcat suggesting an increased turnover number in the presence of urea and GuHCl. Fluorescence studies showed changes in the microenvironment of the protein. The increase in the activity of this intermediate state was due to conformational changes accompanied by increased flexibility at the active site.  相似文献   

20.
Nitric oxide (NO) has a critical role in several physiological and pathophysiological processes. In this paper, the reactions of the nitrosyl complexes of [Ru(bpy)2L(NO)]n+ type, where L = SO32− and imidazole and bpy = 2,2′-bipiridine, with cysteine and glutathione were studied. The reactions with cysteine and glutathione occurred through the formation of two sequential intermediates, previously described elsewhere, [Ru(bpy)2L(NOSR)]n+ and [Ru(bpy)2L(NOSR)2] (SR = thiol) leading to the final products [Ru(bpy)2L(H2O)]n+ and free NO. The second order rate constant for the second step of this reaction was calculated for cysteine k2(SR) = (2.20 ± 0.12) × 109 M− 1 s− 1 and k2(RSH) = (154 ± 2) M− 1 s− 1 for L = SO32− and k2(SR) = (1.30 ± 0.23) × 109 M− 1 s− 1 and k2(RSH) = (0.84 ± 0.02) M− 1 s− 1 for L = imidazole; while for glutathione they were k2(SR) = (6.70 ± 0.32) × 108 M− 1 s− 1 and k2(RSH) = 11.8 ± 0.3 M− 1 s− 1 for L = SO32− and k2(SR) = (2.50 ± 0.36) × 108 M− 1 s− 1 and k2(RSH) = 0.32 ± 0.01 M− 1 s− 1 for L = imidazole. In all reactions it was possible to detect the release of NO from the complexes, which it is remarkably distinct from other ruthenium metallocompounds described elsewhere with just N2O production. These results shine light on the possible key role of NO release mediated by physiological thiols in reaction with these metallonitrosyl ruthenium complexes.  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