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.
Type E botulinum neurotoxin is produced byClostridium botulinum along with a neurotoxin binding protein which helps protect the neurotoxin from adversepH, temperature, and proteolytic conditions. The neurotoxin binding protein has been purified as a 118-kDa protein. Secondary structure content of the neurotoxin binding protein as revealed by far-UV circular dichroism spectroscopy was 19% α-helix, 50%β-sheets, 28% random coils, and 3%β-turns. This compared to 22% α-helix, 44%β-sheets, 34% random coils, and noβ-turns of the type E botulinum neurotoxin. The complex of the two proteins revealed 25%α-helix, 45%β-sheets, 27% random coils, and 3%β-turns, suggesting a significant alteration at least in theα-helical folding of the two proteins upon their interaction. Tyrosine topography is altered considerably (28%) when the neurotoxin and its binding protein are separated, indicating strong interaction between the two proteins. Gel filtration results suggested that type E neurotoxin binding protein clearly complexes with type E neurotoxin. The interaction is favored at lowpH as indicated by an initial binding rate of 8.4 min?1 atpH 5.7 compared to 4.0 min?1 atpH 7.5 as determined using a fiber optic-based biosensor. The neurotoxin and its binding protein apparently are of equivalent antigenicity, as both reacted equally on enzyme-linked immunosorbent assay to polyclonal antibodies raised against the toxoid of their complex.  相似文献   

2.
Lai HT  Lin JS  Chien YH 《Bioresource technology》2011,102(9):5425-5430
This study investigated the effects of light (visible light - 5800 lux, 24 h) or dark regime and aerobic or anaerobic condition on the decay of added oxolinic acid (OA) at 5, 10 and 20 mg L−1 in eel pond sediment. An asymptotic decaying exponential model Ct = Cmin + Co × exp (−k × t) was used to facilitate quantitative approach to OA transformation, where Ct is the concentration of OA after t days, Cmin the estimated level-off concentration of OA residue, Co the concentration of added OA and k the decaying coefficient. OA decayed faster under light (Cmin = 4.6 mg L−1) than under dark (Cmin = 7.8 mg L−1) and also decayed faster under aerobic (Cmin = 4.0 mg L−1) than under anaerobic condition (Cmin = 8.5 mg L−1). Cmin increased with Co. Sundrying and tilling eel pond bottom should be able to reduce OA residue significantly.  相似文献   

3.
Trichoderma asperellum produces two extracellular 1,3-β-d-glucanase upon induction with cell walls from Rhizoctonia solani. A minor 1,3-β-d-glucanase was purified to homogeneity by ion exchange chromatography on Q-Sepharose and gel filtration on Sephacryl S-100. A typical procedure provided 13.8-fold purification with 70% yield. SDS-PAGE of the purified enzyme showed a single protein band of molecular weight 27 kDa. The enzyme exhibited optimum catalytic activity at pH 3.6 and 45 °C. It was thermostable at 40 °C, and retained 75% activity after 60 min at 45 °C. The Km and Vmax values for 1,3-β-d-glucanase, using laminarin as substrate, were 0.323 mg ml−1 and 0.315 U min−1, respectively. The enzyme was strongly inhibited by Hg2+ and SDS. The enzyme was only active toward glucans containing β-1,3-linkages. Peptide sequences showed similarity with two endo-1,3(4)-β-d-glucanases from Aspergillus fumigatus Af293when compared against GenBank non-redundant database.  相似文献   

4.
Human xylosyltransferases I and II (XylT-I, XylT-II) are key enzymes in glycosaminoglycan biosynthesis. Knowledge about the in vivo molecular weight, oligomeric state or turnover number are essential characteristics which have been addressed in this study. XylT-II was purified from Pichia pastoris by fractionated ammonium sulfate precipitation, heparin affinity and ion exchange chromatography. XylT-II was purified over 7000-fold with a final yield of 2.6%. By utilizing mass spectra analysis we can prove its first in-gel detection showing a migration pattern behavior that confirms its in silico molecular weight of 95.8 kDa. We could determine a turnover number of 2.18 min−1 or one transferred xylose molecule per one XylT-II molecule each 27.5 s. The kcat/KM ratio was 0.357 min−1 μM−1 for XylT-II using the bikunin-homologous acceptor Bio-QEEEGSGGGQKK-F. The comparison to XylT-I derived from the same organism revealed a 2.4-fold higher catalytic efficiency (0.870 min−1 μM−1) for XylT-I.  相似文献   

5.
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   

6.
The joint substitution of three active-site residues in Escherichia colil-aspartate aminotransferase increases the ratio of l-cysteine sulfinate desulfinase to transaminase activity 105-fold. This change in reaction specificity results from combining a tyrosine-shift double mutation (Y214Q/R280Y) with a non-conservative substitution of a substrate-binding residue (I33Q). Tyr214 hydrogen bonds with O3 of the cofactor and is close to Arg374 which binds the α-carboxylate group of the substrate; Arg280 interacts with the distal carboxylate group of the substrate; and Ile33 is part of the hydrophobic patch near the entrance to the active site, presumably participating in the domain closure essential for the transamination reaction. In the triple-mutant enzyme, kcat′ for desulfination of l-cysteine sulfinate increased to 0.5 s− 1 (from 0.05 s− 1 in wild-type enzyme), whereas kcat′ for transamination of the same substrate was reduced from 510 s− 1 to 0.05 s− 1. Similarly, kcat′ for β-decarboxylation of l-aspartate increased from < 0.0001 s− 1 to 0.07 s− 1, whereas kcat′ for transamination was reduced from 530 s− 1 to 0.13 s− 1. l-Aspartate aminotransferase had thus been converted into an l-cysteine sulfinate desulfinase that catalyzes transamination and l-aspartate β-decarboxylation as side reactions. The X-ray structures of the engineered l-cysteine sulfinate desulfinase in its pyridoxal-5′-phosphate and pyridoxamine-5′-phosphate form or liganded with a covalent coenzyme-substrate adduct identified the subtle structural changes that suffice for generating desulfinase activity and concomitantly abolishing transaminase activity toward dicarboxylic amino acids. Apparently, the triple mutation impairs the domain closure thus favoring reprotonation of alternative acceptor sites in coenzyme-substrate intermediates by bulk water.  相似文献   

7.
The heme-based oxygen-sensor enzyme from Escherichia coli (Ec DOS) is a heme-regulated phosphodiesterase with activity on cyclic-di-GMP and is composed of an N-terminal heme-bound sensor domain with the PAS structure and a C-terminal functional domain. The activity of Ec DOS is substantially enhanced by the binding of O2 to the Fe(II)-protoporphyrin IX complex [Fe(II) complex] in the sensor domain. The binding of O2 to the Fe(II) complex changes the structure of the sensor domain, and this altered structure becomes a signal that is transduced to the functional domain to trigger catalysis. The first step in intra-molecular signal transduction is the binding of O2 to the Fe(II) complex, and detailed elucidation of this molecular mechanism is thus worthy of exploration. The X-ray crystal structure reveals that Phe113 is located close to the O2 molecule bound to the Fe(II) complex in the sensor domain. Here, we found that the O2 association rate constants (>200 × 10−3 μM−1 s−1: F113L; 26 × 10−3 μM−1 s−1: F113Y) of the Fe(II) complexes of Phe113 mutants were markedly different from that (51 × 10−3 μM−1 s−1) of the wild-type enzyme, and auto-oxidation rates (0.00068 min−1: F113L; 0.039 min−1: F113Y) of the Phe113 mutants also differed greatly from that (0.0062 min−1) of the wild-type enzyme. We thus suggest that Phe113, residing near the O2 molecule, has a critical role in optimizing the Fe(II)-O2 complex for effective regulation of catalysis by the oxygen-sensor enzyme. Interactions of CO and cyanide anion with the mutant proteins were also studied.  相似文献   

8.
ADP-ribosyl cyclase and NAD+ glycohydrolase (CD38, E.C.3.2.2.5) efficiently catalyze the exchange of the nicotinamidyl moiety of NAD+, nicotinamide adenine dinucleotide phosphate (NADP+) or nicotinamide mononucleotide (NMN+) with an alternative base. 4′-Pyridinyl drugs (amrinone, milrinone, dismerinone and pinacidil) were efficient alternative substrates (kcat/KM = 0.9-10 μM−1 s−1) in the exchange reaction with ADP-ribosyl cyclase. When CD38 was used as a catalyst the kcat/KM values for the exchange reaction were reduced two or more orders of magnitude (0.015-0.15 μM−1 s−1). The products of this reaction were novel dinucleotides. The values of the equilibrium constants for dinucleotide formation were determined for several drugs. These enzymes also efficiently catalyze the formation of novel mononucleotides in an exchange reaction with NMN+, kcat/KM = 0.05-0.4 μM−1 s−1. The kcat/KM values for the exchange reaction with NMN+ were generally similar (0.04-0.12 μM−1 s−1) with CD38 and ADP-ribosyl cyclase as catalysts. Several novel heterocyclic alternative substrates were identified as 2-isoquinolines, 1,6-naphthyridines and tricyclic bases. The kcat/KM values for the exchange reaction with these substrates varied over five orders of magnitude and approached the limit of diffusion with 1,6-naphthyridines. The exchange reaction could be used to synthesize novel mononucleotides or to identify novel reversible inhibitors of CD38.  相似文献   

9.
Several RNA-cleaving deoxyribozymes (DNAzymes) have been reported for efficient cleavage of purine-containing junctions, but none is able to efficiently cleave pyrimidine-pyrimidine (Pyr-Pyr) junctions. We hypothesize that a stronger Pyr-Pyr cleavage activity requires larger DNAzymes with complex structures that are difficult to isolate directly from a DNA library; one possible way to obtain such DNAzymes is to optimize DNA sequences with weak activities. To test this, we carried out an in vitro selection study to derive DNAzymes capable of cleaving an rC-T junction in a chimeric DNA/RNA substrate from DNA libraries constructed through chemical mutagenesis of five previous DNAzymes with a kobs of ∼ 0.001 min− 1 for the rC-T junction. After several rounds of selective amplification, DNAzyme descendants with a kobs of ∼ 0.1 min− 1 were obtained from a DNAzyme pool. The most efficient motif, denoted “CT10-3.29,” was found to have a catalytic core of ∼ 50 nt, larger than other known RNA-cleaving DNAzymes, and its secondary structure contains five short duplexes confined by a four-way junction. Several variants of CT10-3.29 exhibit a kobs of 0.3-1.4 min− 1 against the rC-T junction. CT10-3.29 also shows strong activity (kobs  > 0.1 min− 1) for rU-A and rU-T junctions, medium activity (> 0.01 min− 1) for rC-A and rA-T junctions, and weak activity (> 0.001 min− 1) for rA-A, rG-T, and rG-A junctions. Interestingly, a single-point mutation within the catalytic core of CT10-3.29 altered the pattern of junction specificity with a significantly decreased ability to cleave rC-T and rC-A junctions and a substantially increased ability to cleave rA-A, rA-T, rG-A, rG-T, rU-A, and rU-T junctions. This observation illustrates the intricacy and plasticity of this RNA-cleaving DNAzyme in dinucleotide junction selectivity. The current study shows that it is feasible to derive efficient DNAzymes for a difficult chemical task and reveals that DNAzymes require more complex structural solutions for such a task.  相似文献   

10.
The bindings of biogenic polyamines spermine (spm), spermidine (spmd) and synthetic polyamines 3,7,11,15-tetrazaheptadecane·4HCl (BE-333) and 3,7,11,15,19-pentazahenicosane·5HCl (BE-3333) with β-lactoglobulin (β-LG) were determined in aqueous solution. FTIR, UV-vis, CD and fluorescence spectroscopic methods as well as molecular modeling were used to determine the polyamine binding sites and the effect of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind β-LG via both hydrophilic and hydrophobic contacts. Stronger polyamine-protein complexes formed with synthetic polyamines than biogenic polyamines, with overall binding constants of Kspm-β-LG = 3.2(±0.6) × 104 M−1, Kspmd-β-LG = 1.8(±0.5) × 104 M−1, KBE-333-β-LG = 5.8(±0.3) × 104 M−1 and KBE-3333-β-LG = 6.2(±0.05) × 104 M−1. Molecular modeling showed the participation of several amino acids in the polyamine complexes with the following order of polyamine-protein binding affinity: BE-3333 > BE-333 > spermine > spermidine, which correlates with their positively charged amino group content. Alteration of protein conformation was observed with a reduction of β-sheet from 57% (free protein) to 55-51%, and a major increase of turn structure from 13% (free protein) to ∼21% in the polyamine-β-LG complexes, indicating a partial protein unfolding.  相似文献   

11.
In order for cryopreservation to become a practical tool for aquaculture, optimized protocols must be developed for each species and cell type. Knowledge of a cell’s osmotic tolerance and membrane permeability characteristics can assist in optimized protocol development. In this study, these characteristics were determined for Pacific oyster oocytes and modified methods for loading and unloading ethylene glycol (EG) were tested. Oocytes were found to behave as ideal osmometers and their osmotically inactive fraction (Vb) was calculated to be 0.48. Oocytes exposed to NaCl solutions of 0.6 to 2.3 Osm fertilized at rates equivalent to oocytes left in seawater. This corresponds to volume changes of +27.3 and −38.1 ± 1.2%. The permeability of the oocytes to water (Lp) was determined to be 3.8 ± 0.4 × 10−2, 5.7 ± 0.8 × 10−2, and 13.2 ± 1.3 × 10−2 μm min−1 atm−1, when measured at temperatures of 5, 10 and 20 °C. The respective EG permeability values (Ps) were 9.5 ± 0.1 × 10−5, 14.6 ± 1.2 × 10−5, and 41.7 ± 2.4 × 10−5 cm min−1. The activation energies for Lp and Ps were determined to be 14.5 and 17.5 kcal mol−1, respectively. Different models for EG loading and unloading from oocytes were developed and tested. Post-thaw fertilization did not differ significantly between a published step addition method and single step addition at 20 °C. This represents a considerable reduction in handling. The results of this study demonstrate that the cryobiological characteristics of a given cell type should be taken into account when developing cryopreservation methods.  相似文献   

12.
Metagenomic resources representing ruminal bacteria were screened for novel exocellulases using a robotic, high-throughput screening system, the novel CelEx-BR12 gene was identified and the predicted CelEx-BR12 protein was characterized. The CelEx-BR12 gene had an open reading frame (ORF) of 1140 base pairs that encoded a 380-amino-acid-protein with a predicted molecular mass of 41.8 kDa. The amino acid sequence was 83% identical to that of a family 5 glycosyl hydrolase from Prevotella ruminicola 23. Codon-optimized CelEx-BR12 was overexpressed in Escherichia coli and purified using Ni–NTA affinity chromatography. The Michaelis–Menten constant (Km value) and maximal reaction velocity (Vmax values) for exocellulase activity were 12.92 μM and 1.55 × 104 μmol min−1, respectively, and the enzyme was optimally active at pH 5.0 and 37 °C. Multifunctional activities were observed against fluorogenic and natural glycosides, such as 4-methylumbelliferyl-β-d-cellobioside (0.3 U mg−1), CMC (105.9 U mg−1), birch wood xylan (132.3 U mg−1), oat spelt xylan (67.9 U mg−1), and 2-hydroxyethyl-cellulose (26.3 U mg−1). Based on these findings, we believe that CelEx-BR12 is an efficient multifunctional enzyme as endocellulase/exocellulase/xylanase activities that may prove useful for biotechnological applications.  相似文献   

13.
Oscar Goñi 《Phytochemistry》2011,72(9):844-854
A 1,3-β-glucanase with potent cryoprotective activity was purified to homogeneity from the mesocarp of CO2-treated cherimoya fruit (Annona cherimola Mill.) stored at low temperature using anion exchange and chromatofocusing chromatography. This protein was characterized as a glycosylated endo-1,3-β-glucanase with a Mr of 22.07 kDa and a pI of 5.25. The hydrolase was active and stable in a broad acidic pH range and it exhibited maximum activity at pH 5.0. It had a low optimum temperature of 35 °C and it retained 40% maximum activity at 5 °C. The purified 1,3-β-glucanase was relatively heat unstable and its activity declined progressively at temperatures above 50 °C. Kinetic studies revealed low kcat (3.10 ± 0.04 s−1) and Km (0.32 ± 0.03 mg ml−1) values, reflecting the intermediate efficiency of the protein in hydrolyzing laminarin. Moreover, a thermodynamic characterization revealed that the purified enzyme displayed a high kcat at both 37 and 5 °C, and a low Ea (6.99 kJ mol−1) within this range of temperatures. In vitro functional studies indicated that the purified 1,3-β-glucanase had no inhibitory effects on Botrytis cinerea hyphal growth and no antifreeze activity, as determined by thermal hysteresis analysis using differential scanning calorimetry. However, a strong cryoprotective activity was observed against freeze-thaw inactivation of lactate dehydrogenase. Indeed, the PD50 was 8.7 μg ml−1 (394 nM), 9.2-fold higher (3.1 on a molar basis) than that of the cryoprotective protein BSA. Together with the observed accumulation of glycine-betaine in CO2-treated cherimoya tissues, these results suggest that 1,3-β-glucanase could be functionally implicated in low temperature-defense mechanism activated by CO2.  相似文献   

14.
In the present work, we demonstrate that adenine reduced Na+-ATPase activity in isolated basolateral membrane (BLM) of proximal tubule in a dose-dependent manner. Adenine metabolism was ruled out by TLC analysis of the potential [3H]adenine derived-metabolites. Specific binding of [3H]adenine to isolated BLM was observed in a dose-dependent manner with Kd and Bmax of 242.6 ± 27.6 nM and 2749.9 ± 104.9 fmol mg−1, respectively. Adenine increased the [35S]GTPγS specific binding and it was completely abolished by 10−6 M GDPβS (G protein inhibitor) but it was not modified by DPCPX, DMPX and MRS1523, selective antagonists for A1, A2 and A3 receptors, respectively. Furthermore, the inhibitory effect of adenine on the Na+-ATPase activity was blocked by 10−6 M GDPβS, 1 μg/ml pertussis toxin (Gi protein inhibitor), 10−6 M foskolin (adenylyl cyclase activator) and 10−8 M cAMP. These data demonstrate that adenine inhibits the proximal tubule Na+-ATPase activity through the Gi protein-coupled receptor.  相似文献   

15.
The electronic and vibrational Raman spectra of octa-substituted (R = -SC10H21) Co- and Cu-porphyrazines are reported in their solid-state, mesophase, and isotropic liquid forms, as well as in THF solution. Their electronic spectra are composed of traditional Soret (CuS10 = 355 nm, CoS10 = 347 nm) and lower energy Q-bands (CuS10 = 669 nm, CoS10 = 639 nm), as well as a weaker, functionality-specific sulfur n → porphyrin π feature (CuS10 = 500 nm; CoS10 = 447 nm). In contrast to the broad Q-band for CoS10 in all three neat phases, the lower energy analogue for CuS10 is markedly sharper in the microcrystalline state, but similarly broadens in the mesophase, indicative of long range macrocycle π-π interactions that persist even into the liquid state. The resonance (λ = 647 nm) and off-resonance (λ = 785 nm) Raman spectra of these materials in each phase exhibit four diagnostic vibrations; the Cα-Nm stretch (∼1540-1553) cm−1, Cβ-Cβ stretch (∼1450 cm−1), Cα-Cβ-Np stretch (∼1300-1315 cm−1), and Cα-Cβ stretch (∼1070 cm−1). For CoS10, these vibrations systematically shift to lower energy upon melting, while those for CuS10 collapse to degenerate sets. The differences in the electronic and vibrational profiles as a function of temperature suggest that the mesophase structure is governed by strong axial Co-S interactions for CoS10 which template macrocycle π-π stacking, while for CuS10 the same contacts exist, but they are phase dependent and markedly weaker. These inter-porphyrazine interactions are, therefore, responsible for the distinct differences in the melting and clearing temperatures of their respective mesophases. Finally, based on these diagnostic spectroscopic signatures, a photo-thermal, phase-switching mechanism is demonstrated with λ = 785 nm excitation at reduced temperatures, leading to the ability to spectrally monitor and phase change with a single photon source.  相似文献   

16.
The Mg2+ ion-assisted activation mechanism of the active site Tyr8 of a human hematopoietic prostaglandin D2 synthase (H-PGDS) was studied by ultraviolet resonance Raman (UVRR) spectroscopy. Addition of Mg2+ to the native H-PGDS at pH 8.0 resulted in the Y8a Raman band of Tyr8 shifting from 1615 cm−1 to 1600 cm−1. This large shift to lower energy of the tyrosine Y8a vibrational mode is caused by the deprotonation of the tyrosine phenol group promoted by binding of Mg2+. Upon subsequent addition of glutathione (GSH), the Mg2+/H-PGDS solution showed the Tyr8 Raman band shifted to 1611 cm−1, which is 11 cm−1 higher than the frequency of the Mg2+ complex of H-PGDS, but 4 cm−1 lower than the Mg2+ free enzyme. These UVRR observations suggest that the deprotonated Tyr8 in the presence of Mg2+ is re-protonated by the abstraction of H+ from the thiol group of GSH, and that the re-protonated Tyr8 species forms a hydrogen bond with the thiolate anion of GSH. Density functional theory calculations on several model complexes of p-cresol were also performed, which suggested that the pKa and vibrational frequencies of the Tyr8 phenol group are affected by the degree and structure of hydration of the Tyr8 residue.  相似文献   

17.
The aggregates of amyloid beta peptides (Aβs) are regarded as one of the main pathological hallmarks of Alzheimer’s disease (AD). An imbalance between the rates of synthesis and clearance of Aβs is considered to be a possible cause for the onset of AD. Dipeptidyl peptidases II and IV (DPPII and DPPIV) are serine proteases removing N-terminal dipeptides from polypeptides and proteins with proline or alanine on the penultimate position. Alanine is an N-terminal penultimate residue in Аβs, and we presumed that DPPII and DPPIV could cleave them. The results of present in vitro research demonstrate for the first time the ability of DPPIV to truncate the commercial Aβ40 and Aβ42 peptides, to hinder the fibril formation by them and to participate in the disaggregation of preformed fibrils of these peptides. The increase of absorbance at 334 nm due to complex formation between primary amines with o-phtalaldehyde was used to show cleaving of Aβ40 and Aβ42. The time-dependent increase of the quantity of primary amines during incubation of peptides in the presence of DPPIV suggested their truncation by DPPIV, but not by DPPII. The parameters of the enzymatic breakdown by DPPIV were determined for Aβ40 (Km = 37.5 μM, kcat/Km = 1.7 × 103 M−1sec−1) and Aβ42 (Km = 138.4 μM, kcat/Km = 1.90 × 102 M−1sec−1). The aggregation-disaggregation of peptides was controlled by visualization on transmission electron microscope and by Thioflavin-T fluorescence on spectrofluorimeter and fluorescent microscope. DPPIV hindered the peptide aggregation/fibrillation during 3-4 days incubation in 20 mM phosphate buffer, pH 7.4, 37 °C by 50–80%. Ovalbumin, BSA and DPPII did not show this effect. In the presence of DPPIV, the preformed fibrils were disaggregated by 30–40%. Conclusion: for the first time it was shown that the Aβ40 and Aβ42 are substrates of DPPIV. DPPIV prohibits the fibrillation of peptides and promotes disaggregation of their preformed aggregates.  相似文献   

18.
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   

19.
The complex formation of europium(III) and curium(III) with urea in aqueous solution has been studied at I = 0.1 M (NaClO4), room temperature and trace metal concentrations in the pH-range of 1-8 at various ligand concentrations using time-resolved laser-fluorescence spectroscopy. While for curium(III) the luminescence maximum is red shifted upon complexation, in case of europium(III) emission wavelengths remain unaltered but a significant change in peak splitting occurs. Both heavy metals form weak complexes of the formulae ML3+ and MLOH2+ with urea. Stability constants were determined to be log β110 = −0.12 ± 0.05 and log β11-1 = −6.86 ± 0.15 for europium(III) and log β110 = −0.28 ± 0.12 and log β11-1 = −7.01 ± 0.15 for curium(III).  相似文献   

20.
A potent serine proteinase inhibitor was isolated and characterized from the seeds of the tropical legume liana, Derris trifoliata (DtTCI) by ammonium sulfate precipitation, ion exchange chromatography and gel filtration chromatography. SDS-PAGE as well as MALDI-TOF analysis showed that DtTCI is a single polypeptide chain with a molecular mass of ∼20 kDa. DtTCI has three isoinhibitors (pI: 4.55, 5.34 and 5.72) and, inhibited both trypsin and chymotrypsin in a 1:1 molar ratio. Both Dixon plots and Lineweaver-Burk double reciprocal plots revealed a competitive inhibition of trypsin and chymotrypsin activity, with inhibition constants (Ki) of 1.7 × 10−10 and 1.25 × 10−10 M, respectively. N-terminal sequence of DtTCI showed over 50% similarity with numerous Kunitz-type inhibitors of the Papilionoideae subfamily. High pH amplitude and broad temperature optima were noted for DtTCI, and time course experiments indicated a gradual loss in inhibitory potency on treatment with dithiothreitol (DTT). Circular Dichroism (CD) spectrum of native DtTCI revealed an unordered structure whereas exposure to thermal-pH extremes, DTT and guanidine hydrochloride (Gdn HCl) suggested that an abundance of β-sheets along with intramolecular disulfide bonds provide conformational stability to the active site of DtTCI, and that severity of denaturants cause structural modifications promoting inhibitory inactivity. Antimalarial studies of DtTCI indicate it to be a potent antiparasitic agent.  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