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 substrate-inhibitory analysis has shown that single “atypical” cholinesterase (ChE) presents in tissues of freshwater oligochaete Lumbriculus variegatus (O.F. M?ller). This enzyme differs both from “typical” acetylcholinesterase (AChE) and butyrylcholinesterase (BuChE). Specific activity of oligochaete ChE ranges 55–100 μmol ATCh g−1 tissue min−1 or 0.7–1 μmol ATCh mg−1 protein min−1, ratio of maximal rates (V) of substrate hydrolysises is 100:72:71:83 for acetyl-, propionyl-, butyryl- and acetyl-β-metylthiocholine respectively. Values of Michaelis constant (Km) for these substrates are (1.9–2.5) × 10−4 M. The bimolecular enzyme inhibition rate constants (kII) for organophosphorus inhibitors paraoxon, DDVP, and iso-OMPA are 107, 106 и 103 mol−1 | min−1. ATCh and BuTCh exhibit the effect of substrate inhibition of ChE activity, while PrTCh and MeTCh do not.  相似文献   

2.
Biochemical studies to elucidate the structural basis for xyloglucan specificity among GH12 xyloglucanases are lacking. Accordingly, the substrate specificity of a GH12 xyloglucanase from Aspergillus niger (AnXEG12A) was investigated using pea xyloglucan and 12 xylogluco-oligosaccharides, and data were compared to a structural model of the enzyme. The specific activity of AnXEG12A with pea xyloglucan was 113 μmol min−1 mg−1, and apparent kcat and Km values were 49 s−1 and 0.54 mg mL−1, respectively. These values are similar to previously published results using xyloglucan from tamarind seed, and suggest that substrate fucosylation does not affect the specific activity of this enzyme. AnXEG12A preferred xylogluco-oligosaccharides containing more than six glucose units, and with xylose substitution at the −3 and +1 subsites. The specific activities of AnXEG12A on 100 μM XXXGXXXG and 100 μM XLLGXLLG were 60 ± 4 and 72 ± 9 μmol min−1 mg−1, respectively. AnXEG12A did not hydrolyze XXXXXXXG, consistent with other data that demonstrate the requirement for an unbranched glucose residue for hydrolysis by this enzyme.  相似文献   

3.
Saccharomyces cerevisiae Atm1p has been cloned, over-expressed and purified from a yeast expression system. The sequence includes both the soluble ATPase and transmembrane-spanning domains. With the introduction of an N-terminal Kozak sequence and a C-terminal (His)6-tag, a yield of 1 mg of Atm1p was obtained from 3 g wet yeast cells, which is comparable to other membrane-associated proteins isolated from eukaryotic expression systems. The ATPase activity of Atm1p is sensitive to sodium vanadate, a P-type ATPase inhibitor, with an IC50 of 4 μM. MgADP is a product inhibitor for Atm1p with an IC50 of 0.9 mM. The Michaelis–Menten constants Vmax, KM and kcat of Atm1p were measured as 8.7 ± 0.3 μM/min, 107 ± 16 μM and 1.24 ± 0.06 min− 1, respectively. A plot of ATPase activity versus concentration of Atm1p exhibits a nonlinear relationship, suggesting an allosteric response and an important role for the transmembrane domain in mediating both ATP hydrolysis and MgADP release. The metal dependence of Atm1p ATPase activity demonstrated a reactivity order of Mg2+ > Mn2+ > Co2+, while each divalent ion was found to be inhibitory at higher concentrations. The activation and inhibitory effect of phospholipids suggest that formation of a lipid–micelle complex is important for enzymatic activity and stability. Structural analysis of Atm1p by CD spectroscopy suggested a similarity of secondary structure to that found for other members of this ABC protein family.  相似文献   

4.
5.
Kinetics of the reaction of octacarbonyl dicobalt with ethyl diazoacetate leading to [μ2-{ethoxycarbonyl(methylene)}-μ2-(carbonyl)-bis(tricarbonyl-cobalt)] (Co-Co) (1), dinitrogen, and carbon monoxide were investigated at 10 °C in heptane solution. The initial rate of the reaction was measured by following both the gas evolution and the decrease of the octacarbonyl dicobalt concentration. The rate is first order with respect to octacarbonyl dicobalt and a complex order with respect to ethyl diazoacetate and carbon monoxide depending on the ratio of their concentrations. This is in accord with the formation of a heptacarbonyl dicobalt reactive intermediate (k1 (10 °C) = (1.22 ± 0.06) × 10−3 s−1) for which carbon monoxide and ethyl diazoacetate compete (k−1/k2 (10 °C) = 1.34 ± 0.07).  相似文献   

6.
Techniques utilizing β-glucuronidase (GUS) activity as an indicator of Escherichia coli (E. coli) presence use labeled glucuronides to produce optical signals. Carboxyumbelliferyl-β-d-glucuronide (CUGlcU) is a fluorescent labeled glucuronide that is soluble and highly fluorescent at natural water pHs and temperatures and, therefore, may be an ideal reagent for use in an in situ optical sensor. This paper reports for the first time the Michaelis-Menten kinetic parameters for the binding of E. coli GUS with CUGlcU as Km = 910 μM, Vmax = 41.0 μM min−1, Vmax/Km 45.0 μmol L−1 min−1, the optimal pH as 6.5 ± 1.0, optimal temperature as 38 °C, and the Gibb's free energy of activation as 61.40 kJ mol−1. Additionally, it was found CUGlcU hydrolysis is not significantly affected by heavy solvents suggesting proton transfer and solvent addition that occur during hydrolysis are not limiting steps. Comparison studies were made with the more common fluorescent molecule methylumbelliferyl-β-d-glucuronide (MUGlcU). Experiments showed GUS preferentially binds to MUGlcU in comparison to CUGlcU. CUGlcU was also demonstrated in a prototype optical sensor for the detection of E. coli. Initial bench testing of the sensor produced detection of low concentrations of E. coli (1.00 × 103 CFU/100 mL) in 230 ± 15.1 min and high concentrations (1.05 × 105 CFU/100 mL) in 8.00 ± 1.01 min.  相似文献   

7.
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   

8.
Thirteen Psychotria alkaloids were evaluated regarding their interactions with acetylcholinesterase (AChE), butyrylcholinesterase (BChE), and monoamine oxidases A and B (MAO-A and MAO-B), which are enzymatic targets related with neurodegenerative diseases. Two quaternary β-carboline alkaloids, prunifoleine and 14-oxoprunifoleine, inhibited AChE, BChE and MAO-A with IC50 values corresponding to 10 and 3.39 μM for AChE, 100 and 11 μM for BChE, and 7.41 and 6.92 μM for MAO-A, respectively. Both compounds seem to behave as noncompetitive AChE inhibitors and time-dependent MAO-A inhibitors. In addition, the monoterpene indole alkaloids (MIAs) angustine, vallesiachotamine lactone, E-vallesiachotamine and Z-vallesiachotamine inhibited BChE and MAO-A with IC50 values ranging from 3.47 to 14 μM for BChE inhibition and from 0.85 to 2.14 μM for MAO-A inhibition. Among the tested MIAs, angustine is able to inhibit MAO-A in a reversible and competitive way while the three vallesiachotamine-like alkaloids display a time-dependent inhibition on this target. Docking calculations were performed in order to understand the binding mode between the most active ligands and the selected targets. Taken together, our findings established molecular details of AChE, BChE and MAO-A inhibition by quaternary β-carboline alkaloids and MIAs from Psychotria, suggesting these secondary metabolites are scaffolds for the development of multifunctional compounds against neurodegeneration.  相似文献   

9.
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   

10.
Human serum albumin (HSA) participates to heme scavenging, in turn HSA-heme binds gaseous diatomic ligands at the heme-Fe-atom. Here, the effect of abacavir and warfarin on denitrosylation kinetics of HSA-heme-Fe(II)-NO (i.e., koff) is reported. In the absence of drugs, the value of koff is (1.3 ± 0.2) × 10−4 s−1. Abacavir and warfarin facilitate NO dissociation from HSA-heme-Fe(II)-NO, the koff value increases to (8.6 ± 0.9) × 10−4 s−1. From the dependence of koff on the drug concentration, values of the dissociation equilibrium constant for the abacavir and warfarin binding to HSA-heme-Fe(II)-NO (i.e., K = (1.2 ± 0.2) × 10−3 M and (6.2 ± 0.7) × 10−5 M, respectively) were determined. The increase of koff values reflects the stabilization of the basic form of HSA-heme-Fe by ligands (e.g., abacavir and warfarin) that bind to Sudlow’s site I. This event parallels the stabilization of the six-coordinate derivative of the HSA-heme-Fe(II)-NO atom. Present data highlight the allosteric modulation of HSA-heme-Fe(II) reactivity by heterotropic effectors.  相似文献   

11.
The effect of Zn2+ on three major peptidase activities of the 20S proteasome purified from Xenopus oocytes was kinetically investigated. An extremely low concentration of Zn2+ (μM range) strongly inhibited the trypsin-like activity of the 20S proteasome which was fully recoverable by the addition of EDTA. The concentration of Zn2+ for half-maximum inhibition (K0.5) was 0.60 μM which was at least 10 times lower than that of any other divalent cation tested and essentially the same as for proteasomes purified from various other organisms indicating that the inhibition is highly Zn2+-specific, reversible, and common to the proteasome regardless of its source. Zn2+ at concentrations below 100 μM instantaneously activated the chymotrypsin-like and PGPH activities, and the Zn2+ concentration for half-maximum activation was found to be 42-48 μM. These activities were time-dependently inactivated by submillimolar concentrations of Zn2+. The inactivation rates were dependent on the concentration of Zn2+ and reached a maximum of 1.60-2.40 min−1 for the three peptidase activities under the conditions used. The Zn2+ concentration for half-maximum inactivation was found to be 0.70-1.23 mM. This time-dependent inactivation was not reversed by the addition of EDTA or DTT and might not be accompanied by the dissociation of subunits of the 20S proteasome indicating that all activities are inactivated by an identical phenomenon. These results reveal the three types of effects of Zn2+ on the 20S proteasome.  相似文献   

12.
Apigenin, a natural flavone, present in many plants sources, induced apoptosis and cell death in lung epithelium cancer (A549) cells with an IC50 value of 93.7 ± 3.7 μM for 48 h treatment. Target identification investigations using A549 cells and also in cell-free system demonstrated that apigenin depolymerized microtubules and inhibited reassembly of cold depolymerized microtubules of A549 cells. Again apigenin inhibited polymerization of purified tubulin with an IC50 value of 79.8 ± 2.4 μM. It bounds to tubulin in cell-free system and quenched the intrinsic fluorescence of tubulin in a concentration- and time-dependent manner. The interaction was temperature-dependent and kinetics of binding was biphasic in nature with binding rate constants of 11.5 × 10−7 M−1 s−1 and 4.0 × 10−9 M−1 s−1 for fast and slow phases at 37 °C, respectively. The stoichiometry of tubulin–apigenin binding was 1:1 and binding the binding constant (Kd) was 6.08 ± 0.096 μM. Interestingly, apigenin showed synergistic anti-cancer effect with another natural anti-tubulin agent curcumin. Apigenin and curcumin synergistically induced cell death and apoptosis and also blocked cell cycle progression at G2/M phase of A549 cells. The synergistic activity of apigenin and curcumin was also apparent from their strong depolymerizing effects on interphase microtubules and inhibitory effect of reassembly of cold depolymerized microtubules when used in combinations, indicating that these ligands bind to tubulin at different sites. In silico modeling suggested apigenin bounds at the interphase of α–β-subunit of tubulin. The binding site is 19 Å in distance from the previously predicted curcumin binding site. Binding studies with purified protein also showed both apigenin and curcumin can simultaneously bind to purified tubulin. Understanding the mechanism of synergistic effect of apigenin and curcumin could be helped to develop anti-cancer combination drugs from cheap and readily available nutraceuticals.  相似文献   

13.
Endocannabinoids appear to be involved in a variety of physiological processes. Lipoxygenase activity has been known to be affected by unsaturated fatty acids or phenolic compounds. In this study, we examined whether endocannabinoids containing both N-acyl group and phenolic group can affect the activity of soybean lipoxygenase (LOX)-1, similar to mammalian 15-lipoxygenase in physicochemical properties. First, N-arachidonoyl dopamine and N-oleoyl dopamine were found to inhibit soybean LOX-1-catalyzed oxygenation of linoleic acid in a non-competitive manner with a Ki value of 3.7 μM and 6.2 μM, respectively. Meanwhile, other endocannabinoids failed to show a remarkable inhibition of soybean LOX-1. Separately, N-arachidonoyl dopamine and N-arachidonoyl serotonin were observed to inactivate soybean LOX-1 with Kin value of 27 μM and 24 μM, respectively, and k3 value of 0.12 min−1 and 0.35 min−1, respectively. Furthermore, such an inactivation was enhanced by ascorbic acid, but suppressed by 13(S)-hydroperoxy-9,11-octadecadienoic acid. Taken together, it is proposed that endocannabinoids containing polyunsaturated acyl moiety and phenolic group may be efficient for the inhibition as well as inactivation of 15-lipoxygenase.  相似文献   

14.
Tan Y  Sun L  Xi Z  Yang GF  Jiang DQ  Yan XP  Yang X  Li HY 《Analytical biochemistry》2008,383(2):200-204
Protoporphyrinogen oxidase (PPO) is a flavin adenine dinucleotide (FAD)-containing enzyme in the tetrapyrrole biosynthetic pathway that leads to the formation of both heme and chlorophylls, which has been identified as one of the most important action targets of commercial herbicides. The literature reports gave different PPO-catalytic kinetic parameters for the substrate protoporphyrinogen IX (Km of 0.1 to 10.4 μM) with different sources of PPO using fluorescent or HPLC methods. Herein we assayed the enzymatic activity of recombinant Bacillus subtilis PPO by using capillary electrophoresis (CE), a method with high separation efficiency, easy automation, and low sample consumption. The Michaelis constant and maximum reaction velocity were determined as 7.0 ± 0.6 μM and 0.38 ± 0.02 μmol min-1 μg−1, respectively. The interaction between PPO and acifluorfen, a commercial PPO-inhibiting herbicide, was measured as the inhibition constant 186.9 ± 9.3 μМ. The relationship between cofactor FAD and PPO activity can also be quantitatively studied by this CE method. The CE method used here should also be a convenient, reliable method for PPO study.  相似文献   

15.
We collected simultaneous dive Time Depth Recorder (TDR) data and video images from free swimming adult female leatherback turtles, Dermochelys coriacea, during the first 24 h after nesting on the beach, in order to determine relationships between dive parameters, activity, overall respiratory frequency and behaviour.We identified three different underwater locomotory activities (subsurface swimming, V-shaped dives and U-shaped dives) from video and TDR data that varied in their mean depth, duration and a number of other parameters. Overall respiratory frequency (overall fR) was significantly different between all locomotory activities, with turtles taking 1.7±0.1 breaths min−1 while subsurface swimming, 0.78 breaths min−1 after V-shaped dives and 0.57 breaths min−1 after U-shaped dives. Descent rates and ascent rates were significantly faster in U-shaped dives (descent 0.19±0.010 m s−1, ascent 0.28±0.015 m s−1) than in V-shaped dives (descent 0.10±0.008 m s−1, ascent 0.12±0.012 m s−1). Flipper stroke rates were significantly lower during the bottom component of U-shaped dives (0.18±0.02 strokes s−1) than during the descent (0.29±0.03 strokes s−1) or ascent (0.29±0.03 strokes s−1). From overall fR and flipper stroke rate data, we inferred that turtles used less energy during U-shaped dives than the other activity types. We recorded interactions between male turtles and the study females that lasted up to 11 min, during which male turtles displayed the characteristic courtship behaviour of sea turtles. It appeared that females attempted to avoid males by aborting ascent and extending dive duration to swim to the sea floor when males were present.  相似文献   

16.
In this article, we present a new, easy-to-implement assay for methionine γ-lyase (MGL)-catalyzed γ-elimination reactions of l-methionine and its analogues that produce α-ketobutyrate (α-KB) as product. The assay employs ultraviolet–visible (UV–Vis) spectrophotometry to continuously monitor the rate of formation of α-KB by its absorbance at 315 nm. We also employ a nonlinear data analysis method that obviates the need for an “initial slope” determination, which can introduce errors when the progress curves are nonlinear. The spectrophotometric assay is validated through product analysis by 1H NMR (nuclear magnetic resonance), which showed that under the conditions of study l-methionine (l-met) and l-methionine sulfone (l-met sulfone) substrates were converted to α-KB product with greater than 99% yield. Using this assay method, we determined for the first time the Michaelis–Menten parameters for a recombinant form of MGL from Porphyromonas gingivalis, obtaining respective kcat and Km values of 328 ± 8 min−1 and 1.2 ± 0.1 mM for l-met γ-elimination and 2048 ± 59 min−1 and 38 ± 2 mM for l-met sulfone γ-elimination reactions. We envisage that this assay method will be useful for determining the activity of MGL γ-elimination reactions that produce α-KB as the end product.  相似文献   

17.
The complex formation of curium(III) with adenosine 5′-triphosphate (ATP) was determined by time-resolved laser-induced fluorescence spectroscopy (TRLFS). The interaction between soluble species of curium(III) with ATP was studied at trace Cm(III) concentrations (3 × 10−7 M). The concentrations of ATP were varied between 6.0 × 10−7 and 1.5 × 10−4 M in the pH range of 1.5-7.0 using 0.154 M NaCl as background electrolyte.Three Cm-ATP species, MpHqLr, could be identified from the fluorescence emission spectra: (i) CmH2ATP+ with a peak maximum at 598.6 nm, (ii) CmHATP with a peak maximum at 600.3 nm, and (iii) CmATP with a peak maximum at 601.0 nm. The formation constants of these complexes were calculated from TRLFS measurements to be log β121 = 16.86 ± 0.09, log β111 = 13.23 ± 0.10, and log β101 = 8.19 ± 0.16. The hydrated Cm-ATP species showed fluorescence lifetimes between 88 and 96 μs; whereas the CmATP complex has a significantly longer fluorescence lifetime of 187 ± 7 μs.  相似文献   

18.
Investigation of the complex formation and electron transfer kinetics between P450 BMP and flavodoxin was carried out following the suggested involvement of flavodoxin in modulating the electron transfer to BMP in artificial redox chains bound to an electrode surface. While electron transfer measurements show the formation of a tightly bound complex, the NMR data indicate the formation of shortly lived complexes. The measured kobs ranged from 24.2 s− 1 to 44.1 s− 1 with kon ranging from 0.07 × 106 to 1.1 × 106 s− 1M− 1 and Kd ranging from 300 μM to 24 μM in buffers of different ionic strength. This apparent contradiction is due to the existence of two events in the complex formation prior to electron transfer. A stable complex is initially formed. Within such tightly bound complex, flavodoxin rocks rapidly between different positions. The rocking of the bound flavodoxin between several different orientations gives rise to the transient complexes in fast exchange as observed in the NMR experiments. Docking simulations with two different approaches support the theory that there is no highly specific orientation in the complex, but instead one side of the flavodoxin binds the P450 with high overall affinity but with a number of different orientations. The level of functionality of each orientation is dependent on the distance between cofactors, which can vary between 8 and 25 Å, with some of the transient complexes showing distances compatible with the measured electron transfer rate constants.  相似文献   

19.
Phoneutria nigriventer toxin Tx1 (PnTx1, also referred to in the literature as Tx1) exerts inhibitory effect on neuronal (NaV1.2) sodium channels in a way dependent on the holding potential, and competes with μ-conotoxins but not with tetrodotoxin for their binding sites. In the present study we investigated the electrophysiological properties of the recombinant toxin (rPnTx1), which has the complete amino acid sequence of the natural toxin with 3 additional residues: AM on the N-terminal and G on the C-terminal. At the concentration of 1.5 μM, the recombinant toxin inhibits Na+ currents of dorsal root ganglia neurons (38.4 ± 6.1% inhibition at −80 mV holding potential) and tetrodotoxin-resistant Na+ currents (26.2 ± 4.9% at the same holding potential). At −50 mV holding potential the inhibition of the total current reached 71.3 ± 2.3% with 1.5 μM rPnTx1. The selectivity of rPnTx1 was investigated on ten different isoforms of voltage-gated sodium channels expressed in Xenopus oocytes. The order of potency for rPnTx1 was: rNaV1.2 > rNaV1.7 ≈ rNaV1.4 ≥ rNaV1.3 > mNaV1.6 ≥ hNaV1.8. No effect was seen on hNaV1.5 and on the arthropods isoforms (DmNaV1, BGNaV1.1a and VdNaV1). The IC50 for NaV1.2 was 33.7 ± 2.9 nM with a maximum inhibition of 83.3 ± 1.9%. The toxin did not alter the voltage-dependence of channel gating and was effective on NaV1.2 channels devoid of inactivation. It was ineffective on neuronal calcium channels. We conclude that rPnTx1 has a promising selectivity, and that it may be a valuable model to achieve pharmacological activities of interest for the treatment of channelopathies and neuropathic pain.  相似文献   

20.
Based on a broad spectrum of biological activities of rhodanines, we synthesized aromatic amides and esters of 2-(4-oxo-2-thioxothiazolidin-3-yl)acetic acid (rhodanine-3-acetic acid) via carbodiimide- or PCl3-mediated coupling. Both esters and amides were investigated for their in vitro inhibitory potency and selectivity against acetylcholinesterase (AChE) from electric eel and butyrylcholinesterase (BChE) from equine serum using Ellman’s spectrophotometric method. The derivatives exhibited mostly a moderate activity against both cholinesterases. IC50 values for AChE were in a closer concentration range of 24.05–86.85 μM when compared to BChE inhibition (7.92–227.19 μM). The esters caused the more efficient inhibition of AChE than amides and parent acid. The esterification and amidation of the rhodanine-3-acetic acid increased inhibition of BChE, even up to 26 times. Derivatives of 4-nitroaniline/phenol showed the activity superior to other substituents (H, Cl, CH3, OCH3, CF3). Rhodanines produced a balanced inhibition of both cholinesterases. Seven derivatives produced the more potent inhibition of AChE than rivastigmine, a clinically used drug; additional three compounds were comparable. Two amides exceeded inhibitory potency of rivastigmine towards BChE. Importantly, this is the first evidence that rhodanine-based compounds are able to inhibit BChE.  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