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.
Zhang M  Zhang L  Cheung PC 《Biopolymers》2003,68(2):150-159
Seven water-insoluble (1 --> 3)-beta-D-glucan fractions TM8-1 to TM8-7 with weight-average molecular mass M(w) ranged from 2.22 to 77.4 x 10(4) obtained from the sclerotia of Pleurotus tuber-regium were carboxymethylated to produce the water-soluble fractions CTM8-1 to CTM8-7 with M(w) ranged from 3.87 to 87.8 x 10(4). The degree of substitution (DS) of CTM8 fractions was analyzed by ir and elemental analysis (EA) to be 0.3-0.68. The M(w) and the intrinsic viscosity [eta] of the CTM8 fractions were measured by size-exclusion chromatography combined with multiangle laser light scattering (SEC-MALLS), MALLS, and viscometry in phosphate buffer solution (PBS) at 37 degrees C. The dependencies of [eta] and radius of gyration (z) (1/2) on M(w) for the CTM8 samples were found to be [eta] = (8.82 +/- 0.03) x 10(-3) M(w)(0.78 +/- 0.04) (cm(3) g(-1)) and (z) (1/2) = (3.09 +/- 0.05) x 10(-3) M(w)(0.75 +/- 0.06) (nm) in the M(w) range from 3.87 x 10(4) to 53.2 x 10(4). Based on current theories for wormlike chain model, the conformational parameters of the CTM8 were obtained to be 790 (nm(-1)) for M(L), 9.6 (nm) for q, which were higher than those of the native TM8 fractions, suggesting a more extended flexible chain of CTM8 in PBS. On the whole, the CTM8 fractions showed higher antitumor activity than their corresponding TM8 fractions. In view of data from molecular parameters and bioactivity, the antitumor activity of the CTM8 fractions may be correlated to its water solubility and relatively extended chain.  相似文献   

2.
Microcalorimetry has been used to determine enthalpy changes for the hydrolysis of a series of oligosaccharides. High-pressure liquid chromatography was used to determine the extents of reaction and to check for any possible side reactions. The enzyme glucan 1,4-alpha-glucosidase was used to bring about the following hydrolysis reactions: (A) maltose(aq) + H2O(liq) = 2D-glucose(aq); (B) maltotriose(aq) + 2H2O(liq) = 3D-glucose(aq); (C) maltotetraose(aq) + 3H2O(liq) = 4D-glucose(aq); (D) maltopentaose(aq) + 4H2O(liq) = 5D-glucose(aq); (E) maltohexaose(aq) + 5H2O(liq) = 6D-glucose(aq); (F) maltoheptaose(aq) + 6H2O(liq) = 7D-glucose(aq); (G) amylose(aq) + nH2O(liq) = (n + 1) D-glucose(aq); and (H) panose(aq) + 2H2O(liq) = 3D-glucose(aq); (J) isomaltotriose(aq) + 2H2O(liq) = 3D-glucose(aq). The enzyme beta-fructofuranosidase was used for the reactions: (K) raffinose(aq) + H2O(liq) = alpha-D-melibiose(aq) + D-fructose(aq); and (L) stachyose(aq) + H2O(liq) = o-alpha-D-galactopyranosyl-(1----6)- alpha-o-D-galactopyranosyl-(1----6)-alpha-D-glucopyranose + D-fructose(aq). The results of the calorimetric measurements (298.15 K, 0.1 M sodium acetate buffer, pH 4.44-6.00) are: delta H0A = -4.55 +/- 0.10, delta H0B = -9.03 +/- 0.10, delta H0C = -13.79 +/- 0.15, delta H0D = -18.12 +/- 0.10, delta H0E = -22.40 +/- 0.15, delta H0F = -26.81 +/- 0.20, delta H0H = 1.46 +/- 0.40, delta H0J = 11.4 +/- 2.0, delta H0K = -15.25 +/- 0.20, and delta H0L = -14.93 +/- 0.20 kJ mol-1. The enthalpies of hydrolysis of two different samples of amylose were 1062 +/- 20 and 2719 +/- 100 kJ mol-1, respectively.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
The purpose of the present study was to use ratings of perceived exertion (RPE) from the Borg (6-20) and OMNI-Leg (0-10) scales to determine the Physical Working Capacity at the Borg and OMNI thresholds (PWC(BORG) and PWC(OMNI)). PWC(BORG) and PWC(OMNI) were compared with other fatigue thresholds determined from the measurement of heart rate (the Physical Working Capacity at the Heart Rate Threshold: PWC(HRT)), and oxygen consumption (the Physical Working Capacity at the Oxygen Consumption Threshold, PWC(VO2)), as well as the ventilatory threshold (VT). Fifteen men and women volunteers (mean age +/- SD = 22 +/- 1 years) performed an incremental test to exhaustion on an electronically braked ergometer for the determination of VO2 peak and VT. The subjects also performed 4 randomly ordered workbouts to exhaustion at different power outputs (ranging from 60 to 206W) for the determination of PWC(BORG), PWC(OMNI), PWC(HRT), and PWC(VO2). The results indicated that there were no significant mean differences among the fatigue thresholds: PWC(BORG) (mean +/- SD = 133 +/- 37W; 67 +/- 8% of VO2 peak), PWC(OMNI) (137 +/- 44W; 68 +/- 9% of VO2 peak), PWC(HRT) (135 +/- 36W; 68 +/- 8% of VO2 peak), PWC(VO2) (145 +/- 41W; 72 +/- 7% of VO2 peak) and VT (131 +/- 45W; 66 +/- 8% of VO2 peak). The results of this study indicated that the mathematical model used to estimate PWC(HRT) and PWC(VO2) can be applied to ratings of perceived exertion to determine PWC(BORG) and PWC(OMNI) during cycle ergometry. Salient features of the PWC(BORG) and PWC(OMNI) tests are that they are simple to administer and require the use of only an RPE scale, a stopwatch, and a cycle ergometer. Furthermore, the power outputs at the PWC(BORG) and PWC(OMNI) may be useful to estimate the VT noninvasively and without the need for expired gas analysis.  相似文献   

4.
Antitumour glucans (GU) from the fungus Grifora umbellata have been subjected to periodate oxidation, Smith degradation, methylation analysis, and treatment with endo-(1 leads to 6)-beta-D-, endo-(1 leads to 3)-beta-D-, and exo-(1 leads to 3)-beta-D-glucanases, and alpha-amylase; the following structural features were revealed. GU-2 contains a backbone involving (1 leads to 6)-beta- and () leads to 3)-beta linkages, and two kinds of branches involving (1 leads to 6)-beta and (1 leads to 4)-alpha linkages. GU-3 has a (1 leads to 3)-beta-linked backbone and branches involving (1 leads to 6)-beta linkages or (1 leads to 4)-alpha and (1 leads to 6)-beta linkages. GU-4 also contains a (1 leads to 3) beta-D-glucan backbone and a small number of (1 leads to 6)-beta-linked branches. Probable structural units of these glucans are proposed.  相似文献   

5.
Characterization of the GRK2 binding site of Galphaq   总被引:1,自引:0,他引:1  
Heterotrimeric guanine nucleotide-binding proteins (G proteins) transmit signals from membrane bound G protein-coupled receptors (GPCRs) to intracellular effector proteins. The G(q) subfamily of Galpha subunits couples GPCR activation to the enzymatic activity of phospholipase C-beta (PLC-beta). Regulators of G protein signaling (RGS) proteins bind to activated Galpha subunits, including Galpha(q), and regulate Galpha signaling by acting as GTPase activating proteins (GAPs), increasing the rate of the intrinsic GTPase activity, or by acting as effector antagonists for Galpha subunits. GPCR kinases (GRKs) phosphorylate agonist-bound receptors in the first step of receptor desensitization. The amino termini of all GRKs contain an RGS homology (RH) domain, and binding of the GRK2 RH domain to Galpha(q) attenuates PLC-beta activity. The RH domain of GRK2 interacts with Galpha(q/11) through a novel Galpha binding surface termed the "C" site. Here, molecular modeling of the Galpha(q).GRK2 complex and site-directed mutagenesis of Galpha(q) were used to identify residues in Galpha(q) that interact with GRK2. The model identifies Pro(185) in Switch I of Galpha(q) as being at the crux of the interface, and mutation of this residue to lysine disrupts Galpha(q) binding to the GRK2-RH domain. Switch III also appears to play a role in GRK2 binding because the mutations Galpha(q)-V240A, Galpha(q)-D243A, both residues within Switch III, and Galpha(q)-Q152A, a residue that structurally supports Switch III, are defective in binding GRK2. Furthermore, GRK2-mediated inhibition of Galpha(q)-Q152A-R183C-stimulated inositol phosphate release is reduced in comparison to Galpha(q)-R183C. Interestingly, the model also predicts that residues in the helical domain of Galpha(q) interact with GRK2. In fact, the mutants Galpha(q)-K77A, Galpha(q)-L78D, Galpha(q)-Q81A, and Galpha(q)-R92A have reduced binding to the GRK2-RH domain. Finally, although the mutant Galpha(q)-T187K has greatly reduced binding to RGS2 and RGS4, it has little to no effect on binding to GRK2. Thus the RH domain A and C sites for Galpha(q) interaction rely on contacts with distinct regions and different Switch I residues in Galpha(q).  相似文献   

6.
Mechanism of reaction of myeloperoxidase with nitrite   总被引:10,自引:0,他引:10  
Myeloperoxidase (MPO) is a major neutrophil protein and may be involved in the nitration of tyrosine residues observed in a wide range of inflammatory diseases that involve neutrophils and macrophage activation. In order to clarify if nitrite could be a physiological substrate of myeloperoxidase, we investigated the reactions of the ferric enzyme and its redox intermediates, compound I and compound II, with nitrite under pre-steady state conditions by using sequential mixing stopped-flow analysis in the pH range 4-8. At 15 degrees C the rate of formation of the low spin MPO-nitrite complex is (2.5 +/- 0.2) x 10(4) m(-1) s(-1) at pH 7 and (2.2 +/- 0.7) x 10(6) m(-1) s(-1) at pH 5. The dissociation constant of nitrite bound to the native enzyme is 2.3 +/- 0.1 mm at pH 7 and 31.3 +/- 0.5 micrometer at pH 5. Nitrite is oxidized by two one-electron steps in the MPO peroxidase cycle. The second-order rate constant of reduction of compound I to compound II at 15 degrees C is (2.0 +/- 0.2) x 10(6) m(-1) s(-1) at pH 7 and (1.1 +/- 0.2) x 10(7) m(-1) s(-1) at pH 5. The rate constant of reduction of compound II to the ferric native enzyme at 15 degrees C is (5.5 +/- 0.1) x 10(2) m(-1) s(-1) at pH 7 and (8.9 +/- 1.6) x 10(4) m(-1) s(-1) at pH 5. pH dependence studies suggest that both complex formation between the ferric enzyme and nitrite and nitrite oxidation by compounds I and II are controlled by a residue with a pK(a) of (4.3 +/- 0.3). Protonation of this group (which is most likely the distal histidine) is necessary for optimum nitrite binding and oxidation.  相似文献   

7.
The S(3) state of the water-oxidizing complex (WOC) of photosystem II (PSII) is the last state that can be trapped before oxygen evolution occurs at the transient S(4) state. A number of EPR-detectable intermediates are associated with this critical state. The preceding paper examined mainly the decay of S(3) at cryogenic temperatures leading to the formation of a proton-deficient configuration of S(2) termed S(2)'. This second paper examines all intermediates formed by the near-IR light (NIR) excitation of the S(3) state and compares these with the light-excitation products of the S(2)' state. The rather complex set of observations is organized in a comprehensive flowchart, the central part of which is the S(3)...Q(A)(-) state. This state can be converted to various intermediates via two main pathways: (A) Excitation of S(3) by NIR light at temperatures below 77 K results presumably in the formation of an excited S(3) state, S(3), which decays via either of two pathways. Slowly at liquid helium temperatures but much faster at 77 K, S(3) decays to an EPR-silent state, denoted S(3)' ', which by raising the temperature to ca. 190 K converts to a spin configuration of the Mn cluster, characterized by g = 21, 3.7 in perpendicular and g = 23 in parallel mode EPR, denoted S(3)'. Upon further warming to 220 K, S(3)' relaxes to the untreated S(3) state. Below about 77 K and more favorably at liquid helium temperatures, an alternative pathway of S(3) decay via the metallo-radical intermediate S(2)'Z*...Q(A)(-) can be traced. This leads to the metastable state S(2)'Z...Q(A) via charge recombination. S(2)'Z* is characterized by a split-radical signal at g = 2, while all S(2)' transients are characterized by the same g = 5/2.9 (S = (7)/(2)) configuration of the Mn cluster with small modifications, reflecting an influence of the tyr Z oxidation state on the crystal-field symmetry at the Mn cluster. (B) S(2)'...Q(A) can be reached alternatively by the slow charge recombination of S(3) and Q(A)(-) at 77 K. White-light illumination of S(2)'.Q(A) below about 20 K results in charge separation, reforming the intermediate S(2)'Z*...Q(A)(-). Thermally activated branches to the main pathways are also described, e.g., at elevated temperatures tyr Z* reoxidizes S(2)' to the S(3) state. The above observations are discussed in terms of a molecular model of the S(3) state of the OEC. Main aspects of the model are the following. Intermediates, isoelectronic to S(3), are attributed to the NIR-induced translocation of the positive hole to different Mn ligands, or to tyr Z. On the basis of a comparison of the electron-donating efficiency of tyr Z and tyr D at cryogenic temperatures, it is inferred that the Mn cluster acts as the main proton acceptor from tyr Z. Water associated with the Mn cluster is assumed to be in hydrogen-bonding equilibrium with tyr Z, and an array comprising this water and adjacent water (or OH or O) ligands to Mn followed by a sequence of proton acceptors is proposed to act as an efficient proton translocation pathway. Oxidation of the tyrosine by P(680)(+) repels protons to and out from the Mn cluster. This proposed role of tyr Z in the water-splitting process is described as a proton repeller/electron abstractor.  相似文献   

8.
Kopper BJ  Lindroth RL 《Oecologia》2003,134(1):95-103
The purpose of this study was to assess the independent and interactive effects of CO(2), O(3), and plant genotype on the foliar quality of a deciduous tree and the performance of a herbivorous insect. Two trembling aspen (Populus tremuloides Michaux) genotypes differing in response to CO(2) and O(3) were grown at the Aspen FACE (Free Air CO(2) Enrichment) site located in northern Wisconsin, USA. Trees were exposed to one of four atmospheric treatments: ambient air (control), elevated carbon dioxide (+CO(2); 560 microl/l), elevated ozone (+O(3); ambient x1.5), and elevated CO(2)+O(3). We measured the effects of CO(2) and O(3) on aspen phytochemistry and on performance of forest tent caterpillar (Malacosoma disstria Hübner) larvae. CO(2) and O(3) treatments influenced foliar quality for both genotypes, with the most notable effects being that elevated CO(2) reduced nitrogen and increased tremulacin levels, whereas elevated O(3) increased early season nitrogen and reduced tremulacin levels, relative to controls. With respect to insects, the +CO(2) treatment had little or no effect on larval performance. Larval performance improved in the +O(3) treatment, but this response was negated by the addition of elevated CO(2) (i.e., +CO(2)+O(3) treatment). We conclude that tent caterpillars will have the greatest impact on aspen under current CO(2) and high O(3) levels, due to increases in insect performance and decreases in tree growth, whereas tent caterpillars will have the least impact on aspen under high CO(2) and low O(3) levels, due to moderate changes in insect performance and increases in tree growth.  相似文献   

9.
Payne JC  Rous BW  Tenderholt AL  Godwin HA 《Biochemistry》2003,42(48):14214-14224
Zinc binding to the two Cys(4) sites present in the DNA-binding domain (DBD) of nuclear hormone receptor proteins is required for proper folding of the domain and for protein activity. By utilizing Co(2+) as a spectroscopic probe, we have characterized the metal-binding properties of the two Cys(4) structural zinc-binding sites found in the DBD of human estrogen receptor alpha (hERalpha-DBD) and rat glucocorticoid receptor (GR-DBD). The binding affinity of Co(2+) to the two proteins was determined relative to the binding affinity of Co(2+) to the zinc finger consensus peptide, CP-1. Using the known dissociation constant of Co(2+) from CP-1, the dissociation constants of cobalt from hERalpha-DBD were calculated: K(d1)(Co) = 2.2 (+/- 1.0) x 10(-7) M and K(d2)(Co) = 6.1 (+/- 1.5) x 10(-7) M. Similarly, the dissociation constants of Co(2+) from GR-DBD were calculated: K(d1)(Co) = 4.1 (+/- 0.6) x 10(-7) M and K(d2)(Co) = 1.7 (+/- 0.3) x 10(-7) M. Metal-binding studies conducted in which Zn(2+) displaces Co(2+) from the metal-binding sites of hERalpha-DBD and GR-DBD indicate that Zn(2+) binds to each of the Cys(4) metal-binding sites approximately 3 orders of magnitude more tightly than Co(2+) does: the stoichiometric dissociation constants are K(d1)(Zn) = 1 (+/- 1) x 10(-10) M and K(d2)(Zn) = 5 (+/- 1) x 10(-10) M for hERalpha-DBD and K(d1)(Zn) = 2 (+/- 1) x 10(-10) M and K(d2)(Zn) = 3 (+/- 1) x 10(-10) M for GR-DBD. These affinities are comparable to those observed for most other naturally occurring structural zinc-binding sites. In contrast to the recent prediction by Low et. al. that zinc binding in these systems should be cooperative [Low, L. Y., Hernández, H., Robinson, C. V., O'Brien, R., Grossmann, J. G., Ladbury, J. E., and Luisi, B. (2002) J. Mol. Biol. 319, 87-106], these data suggest that the zincs that bind to the two sites in the DBDs of hERalpha-DBD and GR-DBD do not interact.  相似文献   

10.
The inositol lipid and phosphate binding properties and the cellular localization of phospholipase Cdelta(4) (PLCdelta(4)) and its isolated pleckstrin homology (PH) domain were analyzed in comparison with the similar features of the PLCdelta(1) protein. The isolated PH domains of both proteins showed plasma membrane localization when expressed in the form of a green fluorescent protein fusion construct in various cells, although a significantly lower proportion of the PLCdelta(4) PH domain was membrane-bound than in the case of PLCdelta(1)PH-GFP. Both PH domains selectively recognized phosphatidylinositol 4,5-bisphosphate (PI(4,5)P(2)), but a lower binding of PLCdelta(4)PH to lipid vesicles containing PI(4,5)P(2) was observed. Also, higher concentrations of inositol 1,4,5-trisphosphate (Ins(1,4,5)P(3)) were required to displace the PLCdelta(4)PH from the lipid vesicles, and a lower Ins(1,4,5)P(3) affinity of PLCdelta(4)PH was found in direct Ins(1,4,5)P(3) binding assays. In sharp contrast to the localization of its PH domain, the full-length PLCdelta(4) protein localized primarily to intracellular membranes mostly to the endoplasmic reticulum (ER). This ER localization was in striking contrast to the well documented PH domain-dependent plasma membrane localization of PLCdelta(1). A truncated PLCdelta(4) protein lacking the entire PH domain still showed the same ER localization as the full-length protein, indicating that the PH domain is not a critical determinant of the localization of this protein. Most important, the full-length PLCdelta(4) enzyme still showed binding to PI(4,5)P(2)-containing micelles, but Ins(1,4,5)P(3) was significantly less potent in displacing the enzyme from the lipid than with the PLCdelta(1) protein. These data suggest that although structurally related, PLCdelta(1) and PLCdelta(4) are probably differentially regulated in distinct cellular compartments by PI(4,5)P(2) and that the PH domain of PLCdelta(4) does not act as a localization signal.  相似文献   

11.
Several pure poly(I) preparations differed in: (a) their complement fixation reactivity with anti-poly(I) antiserum; (b) their ability to bind to a solid-phase anti-poly(I) antibody-Sepharose column; (c) their ability to inactivate serum complement; and (d) their reactivity with purified antibodies to double-stranded RNA. In particular, poly(I) samples that could induce interferon production differed from non-inducer poly(I)s; the inducers reacted weakly with anti-poly(I) antiserum and were the only ones that reacted with antibodies to double-stranded RNA. One inducer poly(I) did not inactivate complement, and differed from non-inducer poly(I) in quantitative aspects of poly(I) . poly(C) formation with varying amounts of poly(C). An additional type of poly(I) preparation reacted poorly with anti-poly(I) antiserum, did not react with anti-double-stranded-RNA antibodies and failed to induce interferon production. The varying forms of poly(I) were not interconvertible by boiling and rapid chilling. These results indicate that several different stable structural forms of poly(I) may result from a standardized synthetic procedure.  相似文献   

12.
We report the results of NMR studies and computer simulations of potent antagonists reflective of the alpha(IIb)beta(3) receptor-bound conformations. The peptides c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Phe(4)-(15)N-Arg(5)-Cys]-NH(2) (Phe-Arg analog) (Mpa: 3-mercaptopropionic acid) and c[Mpa-(15)N-Arg(1)-(15)N-Gly(2)-(15)N-Asp(3)-(15)N-Asp(4)-(15)N-Val(5)-Cys]-NH(2) (Asp-Val analog) were subjected to (15)N-edited NMR experiments to study the conformations of these peptides in the absence and in the presence of alpha(IIb)beta(3) receptor. The NMR studies of the Phe-Arg analog, a selective alpha(IIb)beta(3) antagonist, resulted in distinctly different experimental data in the presence and absence of the receptor. The computer simulations for this peptide resulted in one large family of structures consistent with the experimental data. This conformation suggests a type I beta-turn spanning residues Arg(1) and Gly(2) when bound to the receptor and we were able to establish a model for the three dimensional arrangement of the pharmacophores. The studies on the Asp-Val analog, an alpha(v)beta(3) antagonist that binds to the alpha(IIb)beta(3) with moderate affinity, resulted in conformations that are not as well defined as those for the Phe-Arg analog but are consistent with the model established for this analog. These results are important for the design of novel alpha(IIb)beta(3) antagonists.  相似文献   

13.
The adsorption of a range of single chain zwitterionic phosphocholine surfactants (C(n)P(m)C) at the air/liquid interface has been studied by a combination of surface tension and neutron reflectivity. The critical micellar concentration (CMC) for C(n)PC (or C(n)P(2)C), where n varied from 12, 14 to 16, was found to be 0.91, 0.14, and 1.2 x 10(-2) mM respectively, and followed the same trend as observed for other zwitterionic and non-ionic surfactants. The area per molecule at the CMC, A(cmc), for C(n)PC was found to remain constant between 50 and 53 A(2), indicating that the increase in the alkyl chain length had little effect on A(cmc) at the interface. The neutron reflection measurement also showed an almost constant layer thickness (tau) of 20+/-2 A from all the alkyl chain deuterated PC surfactants (dC(n)hPC) in null reflecting water (NRW), suggesting that the alkyl chains of the surfactant responded to changes in either chain length or solution concentration by varying their angle of tilt. In contrast, increasing the length of head group linker between P and N atoms in C(12)P(m)C, where m=2, 4, to 6, resulted in a much slower decrease of CMC from 0.91, 0.7, to 0.5 mM, consistent with a different contribution to the free energy of micellization. A(cmc) for C(12)P(m)C did not vary when m was increased from 2 to 4, and this observation together with the thickness of the head group region indicated an almost perpendicular projection of the head group in C(12)P(2)C and C(12)P(4)C. A further increase in m to 6 resulted in an A(cmc) of 70 A(2). This increase in A(cmc) however did not result in any change in either the total layer thickness or the fraction of the head group region submerged in the aqueous subphase, suggesting that the head group in C(12)P(6)C was bent away from the surface normal direction. Both increase in temperature from 25 to 40 degrees C and the addition of 0.1 M NaCl had little effect on the area per molecule or the thickness of C(12)P(m)C surfactant layer, showing that the C(12)P(m)C series behaved like C(n)P(2)C series. The main conclusion from this study is that for all the C(n)P(m)C surfactants studied, change in m or n has little effect on the total thickness, the thickness of the alkyl chain or that of the head group region.  相似文献   

14.
The potential of anaflatoxin B(1) (AnAFB(1)) conjugated to keyhole limpet hemocyanin (KLH) as a vaccine (AnAFB(1)-KLH) in controlling the carry over of the aflatoxin B(1) (AFB(1)) metabolite aflatoxin M(1) (AFM(1)) in cow milk is reported. AFB(1) is the most carcinogenic compound in food and foodstuffs amongst aflatoxins (AFs). AnAFB(1) is AFB(1) chemically modified as AFB(1)-1(O-carboxymethyl) oxime. In comparison to AFB(1), AnAFB(1) has proven to be non-toxic in vitro to human hepatocarcinoma cells and non mutagenic to Salmonella typhimurium strains. AnAFB(1)-KLH was used for immunization of cows proving to induce a long lasting titer of anti-AFB(1) IgG antibodies (Abs) which were cross reactive with AFB(1), AFG(1), and AFG(2). The elicited anti-AFB(1) Abs were able to hinder the secretion of AFM(1) into the milk of cows continuously fed with AFB(1). Vaccination of lactating animals with conjugated AnAFB(1) may represent a solution to the public hazard constituted by milk and cheese contaminated with AFs.  相似文献   

15.
To analyze functional differences in the interactions of the glucagon receptor (GR) with the two predominant splice variants of Galpha(s), GR was covalently linked to the short and the long forms Galpha(s)-S and Galpha(s)-L to produce the fusion proteins GR-Galpha(s)-S and GR-Galpha(s)-L. GR-Galpha(s)-S bound glucagon with an affinity similar to that of GR, while GR-Galpha(s)-L showed a 10-fold higher affinity for glucagon. In the presence of GTPgammaS, GR-Galpha(s)-L reverted to the low affinity glucagon binding conformation. Both GR-Galpha(s)-L and GR-Galpha(s)-S were constitutively active, causing elevated basal levels of cAMP even in the absence of glucagon. A mutant GR that failed to activate G(s) (G23D1R) was fused to Galpha(s)-L. G23D1R-Galpha(s)-L bound glucagon with high affinity, but failed to elevate cAMP levels, suggesting that the mechanisms of GR-mediated Galpha(s)-L activation and Galpha(s)-L-induced high affinity glucagon binding are independent. Both GR-Galpha(s)-S and GR-Galpha(s)-L bound the antagonist desHis(1)[Nle(9),Ala(11),Ala(16)]glucagon amide with affinities similar to GR. The antagonist displayed partial agonist activity with GR-Galpha(s)-L, but not with GR-Galpha(s)-S. Therefore, the partial agonist activity of the antagonist observed in intact cells appears to be due to GRs coupled to Galpha(s)-L. We conclude that Galpha(s)-S and Galpha(s)-L interact differently with GR and that specific coupling of GR to Galpha(s)-L may account for GTP-sensitive high affinity glucagon binding.  相似文献   

16.
Saccharomyces cerevisiae phosphoenolpyruvate (PEP) carboxykinase catalyzes the reversible formation of oxaloacetate and adenosine triphosphate from PEP, adenosine diphosphate and carbon dioxide, and uses Mn(2+) as the activating metal ion. Comparison with the crystalline structure of homologous Escherichia coli PEP carboxykinase [Tari et al. Nature Struct. Biol. 4 (1997) 990-994] shows that Lys(213) is one of the ligands to Mn(2+) at the enzyme active site. Coordination of Mn(2+) to a lysyl residue is infrequent and suggests a low pK(a) value for the epsilon-NH(2) group of Lys(213). In this work, we evaluate the role of neighboring Phe(416) in contributing to provide a low polarity microenvironment suitable to keep the epsilon-NH(2) of Lys(213) in the unprotonated form. Mutation Phe416Tyr shows that the introduction of a hydroxyl group in the lateral chain of the residue produces a substantial loss in the enzyme affinity for Mn(2+), suggesting an increase of the pK(a) of Lys(213). A study of the effect of pH on K(m) for Mn(2+) indicate that the affinity of recombinant wild type enzyme for the metal ion is dependent on deprotonation of a group with pK(a) of 7.1+/-0.2, compatible with the low pK(a) expected for Lys(213). This pK(a) value increases at least 1.5 pH units upon Phe416Tyr mutation, in agreement with the expected effect of an increase in the polarity of Lys(213) microenvironment. Theoretical calculations of the pK(a) of Lys(213) indicate a value of 6.5+/-0.9, and it increases to 8.2+/-1.6 upon Phe416Tyr mutation. Additionally, mutation Phe416Tyr causes a loss of 1.3 kcal mol(-1) in the affinity of the enzyme for PEP, an effect perhaps related to the close proximity of Phe(416) to Arg(70), a residue previously shown to be important for PEP binding.  相似文献   

17.
The active site of human salivary alpha-amylase is composed of tandem subsites (S3, S2, S1, S1',S2', etc.) geometrically complementary to several glucose residues, and the glycosidic linkage of the substrate is split between S1 and S1'. As a matter of convenience, the subsites to which the non-reducing-end part (glycone) and the reducing-end part (aglycone) of the substrate being hydrolyzed are bound are named the glycone-binding site (S3, S2, S1) and the aglycone-binding site (S1', S2'), respectively. The features of the aglycone-binding site of human salivary alpha-amylase were examined by means of transglycosylation reaction using phenyl alpha-maltoside (GG phi: G-G-phi) and its derivatives (GAG phi: G-AG-phi, GCG phi: G-CG-phi, AGG phi: AG-G-phi, and CGG phi: CG-G-phi) in which one of the glucose residues (G) has been converted to 6-amino-6-deoxy-glucose (AG) or glucuronic acid (CG) residue as the acceptor. A fluorogenic derivative of maltotetraose, p-nitrophenyl O-6-deoxy-6-[(2-pyridyl)amino]-alpha-D-glucopyranosyl-(1----4)-O-alpha-D -glucopyranosyl-(1----4)-O-alpha-D-glucopyranosyl-(1----4)-alpha-D- glucopyranosyl-(1----4)-alpha-D-glucopyranoside (FG4P, FG-G-G-G-P), was used as the substrate. HSA catalyzed both hydrolysis of FG4P to FG3 (FG-G-G) and p-nitrophenyl alpha-glucoside (G-P) and transfer of the FG3 residue of FG4P to the acceptors. Transfer to GAG phi occurred more effectively than to GG phi. Transfers to GCG phi and CGG phi were less than to GG phi and very little transfer to AGG phi occurred.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Yamazaki S  Morioka C  Itoh S 《Biochemistry》2004,43(36):11546-11553
Tyrosinase is a copper monooxygenase containing a coupled dinuclear copper active site (type-3 copper), which catalyzes oxygenation of phenols (phenolase activity) as well as dehydrogenation of catechols (catecholase activity) using O(2) as the oxidant. In this study, catalase activity (conversion of H(2)O(2) to (1/2)O(2) and H(2)O) and peroxygenase activity (H(2)O(2)-dependent oxygenation of substrates) of mushroom tyrosinase have been examined kinetically by using amperometric O(2) and H(2)O(2) sensors. The catalase activity has been examined by monitoring the initial rate of O(2) production from H(2)O(2) in the presence of a catalytic amount of tyrosinase in 0.1 M phosphate buffer (pH 7.0) at 25 degrees C under initially anaerobic conditions. It has been found that the catalase activity of mushroom tyrosinase is three-order of magnitude greater than that of mollusk hemocyanin. The higher catalase activity of tyrosinase could be attributed to easier accessibility of H(2)O(2) to the dinuclear copper site of tyrosinase. Mushroom tyrosinase has also been demonstrated for the first time to catalyze oxygenation reaction of phenols with H(2)O(2) (peroxygenase activity). The reaction has been investigated kinetically by monitoring the H(2)O(2) consumption rate in 0.5 M borate buffer (pH 7.0) under aerobic conditions. Similarity of the substituent effects of a series of p-substituted phenols in the peroxygenase reaction with H(2)O(2) to those in the phenolase reaction with O(2) as well as the absence of kinetic deuterium isotope effect with a perdeuterated substrate (p-Cl-C(6)D(4)OH vs p-Cl-C(6)H(4)OH) clearly demonstrated that the oxygenation mechanisms of phenols in both systems are the same, that is, the electrophilic aromatic substitution reaction by a (micro-eta(2):eta(2)-peroxo)dicopper(II) intermediate of oxy-tyrosinase.  相似文献   

19.
We report different analytical methods used to study the effects of 3\'-azido-3\'-deoxythymidine, aspirin, taxol, cisplatin, atrazine, 2,4-dichlorophenoxyacetic, biogenic polyamines, chlorophyll, chlorophyllin, poly(ethylene glycol), vanadyl cation, vanadate anion, cobalt-hexamine cation, and As2O3, on the stability and secondary structure of human serum albumin (HSA) in aqueous solution, using capillary electrophoresis, Fourier transform infrared, ultraviolet visible, and circular dichroism (CD) spectroscopic methods. The concentrations of HSA used were 4% to 2% or 0.6 to 0.3 mM, while different ligand concentrations were 1 microM to 1 mM. Structural data showed drugs are mostly located along the polypeptide chains with both specific and nonspecific interactions. The stability of drug-protein complexes were in the order K(VO(2+)) 1.2 x 10(8) M(-1) > K(AZT) 1.9 x 10(6) M(-)1 > K(PEG) 4.1 x 10(5) M(-1) > K(atrazine) 3.5 x 10(4) M(-1) > K(chlorophyll) 2.9 x 10(4) M(-1) > K2,4-D 2.5 x 10(4) M-1 > K(spermine) 1.7 x 10(4) M(-1) > K(taxol) 1.43 x 10(4) M(-1) > K(Co(3+)) > 1.1 x 10(4) M(-1) > K(aspirin) 1.04 x 10(4)i(-1) > K(chlorophyllin) 7.0 x 10(3) M(-1) > K(VO(3)(-)) 6.0 x 103 M(-1) > K(spermidine) 5.4 x 10(3) M(-1) > K(putrescine) 3.9 x 10(3) M(-1) > K(As(2)O(3)) 2.2 x 10(3) M(-1)> K(cisplatin) 1.2 x 10(2) M(-1). The protein conformation was altered (infrared and CD results) with major reduction of alpha-helix from 60 to 55% (free HSA) to 49 to 40% and increase of beta-structure from 22 to 15% (free HSA) to 33 to 23% in the drug-protein complexes. The alterations of protein secondary structure are attributed to a partial unfolding of HSA on drug complexation.  相似文献   

20.
The effect of phosphorylation on the shape of tyrosine hydroxylase (TH) was studied directly using gel filtration and indirectly using electrospray ionization mass spectrometry. Phosphorylation of Ser(19) and Ser(40) produced a TH molecule with a more open conformation than the non-phosphorylated form. The conformational effect of Ser(19) phosphorylation is less pronounced than that of the Ser(40) phosphorylation. The effect of Ser(19) and Ser(40) phosphorylation appears to be additive. Binding of dopamine produced a more compact form when compared with the non-dopamine-bound TH. The interdependence of Ser(19) and Ser(40) phosphorylation was probed using electrospray ionization mass spectrometry. The rate constants for the phosphorylation of Ser(19) and Ser(40) were determined by electrospray ionization mass spectrometry using a consecutive reaction model. The rate constant for the phosphorylation of Ser(40) is approximately 2- to 3-fold higher if Ser(19) is already phosphorylated. These results suggest that phosphorylation of Ser(19) alters the conformation of tyrosine hydroxylase to allow increased accessibility of Ser(40) to kinases.  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