共查询到20条相似文献,搜索用时 0 毫秒
1.
Glyn Williams Crispin G.S. Eley Geoffrey R. Moore Martin N. Robinson Robert J.P. Williams 《FEBS letters》1982,150(2):293-299
The interaction of horse ferricytochrome c with the reagents [Fe(EDTA)(H2O)]− and [Cr(CN)6]3− were studied at pH 7 and 25°C by 1H-NMR spectroscopy. Two binding regions near to the heme crevice of cytochrome c were identified. Both regions bound both reagents but they exhibited different selectivities.
The relevance of this finding to the electron-transfer function of cytochrome c is discussed. 相似文献
2.
Whittaker SB Czisch M Wechselberger R Kaptein R Hemmings AM James R Kleanthous C Moore GR 《Protein science : a publication of the Protein Society》2000,9(4):713-720
The bacterial toxin colicin E9 is secreted by producing Escherichia coli cells with its 9.5 kDa inhibitor protein Im9 bound tightly to its 14.5 kDa C-terminal DNase domain. Double- and triple-resonance NMR spectra of the isolated DNase domain uniformly labeled with 13C/15N bound to unlabeled Im9 contain more signals than expected for a single DNase conformer, consistent with the bound DNase being present in more than one form. The presence of chemical exchange cross peaks in 750 MHz 15N-1H-15N HSQC-NOESY-HSQC spectra for backbone NH groups of Asp20, Lys21, Trp22, Leu23, Lys69, and Asn70 showed that the bound DNase was in dynamic exchange. The rate of exchange from the major to the minor form was determined to be 1.1 +/- 0.2 s(-1) at 298 K. Previous NMR studies have shown that the free DNase interchanges between two conformers with a forward rate constant of 1.61 +/- 0.11 s(-1) at 288 K, and that the bound Im9 is fixed in one conformation. The NMR studies of the bound DNase show that Im9 binds similarly to both conformers of the DNase and that the buried Trp22 is involved in the dynamic process. For the free DNase, all NH groups within a 9 A radius of any point of the Trp22 ring exhibit heterogeneity suggesting that a rearrangement of the position of this side chain is connected with the conformational interchange. The possible functional significance of this feature of the DNase is discussed. 相似文献
3.
The role of the heme propionate groups in determining the electron transfer and electrostatic properties of myoglobin have been studied by thermodynamic, kinetic, and spectroscopic studies of horse heart myoglobin in which the heme propionate groups are esterified. Spectroelectrochemical analysis has established that the Em,7 of dimethylester heme-substituted Mb (DME-Mb) (Em,7 = 100.2(2) mV vs. NHE (Normal Hydrogen Electrode) (25 °C) is increased 40 mV relative to that of the native protein with ΔH° = −12.9(2) kcal/mol and ΔS° = −51.0(8) cal/mol/deg (pH 7.0, μ = 0.1 M (phosphate)). The second order rate constant for reduction of DME-metMb by Fe(EDTA)2− is increased > 400-fold relative to that for reduction of native metMb to a value of 1.34(2) × 103 M−1 s−1 with ΔS‡ = −13(1) cal/mol/deg and ΔH‡ = 9.2(3) (pH 7.0, μ = 0.1 M (phosphate)). Analysis of the pH dependences of the reduction potential and rate constant for reduction by Fe(EDTA)2− demonstrates that heme propionate esterification introduces significant changes into the electrostatic interactions in myoglobin. These changes are also manifested by differences in the pH dependences of the 1H NMR spectra of native and DME-metMb that reveal shifts in pKa values for specific His residues as the result of heme propionate esterification. In sum, the current results establish that heme propionate esterification not only affects the electron transfer properties of myoglobin but also influences the titration behavior of specific His residues. 相似文献
4.
Tzvetanova E Nenkova G Georgieva A Alexandrova A Girchev R Kirkova M 《Cell biochemistry and function》2011,29(2):135-141
The in vivo effects of nociceptin (N/OFQ(1–13)NH2) and its structural analogues ([Dab9]N/OFQ(1–13)NH2, [Dap9]N/OFQ(1–13)NH2 and [Cav9]N/OFQ(1–13)NH2) on the levels of lipid peroxidation and cell antioxidants (enzyme and non‐enzyme) in brain of control and kainic acid (KA)‐treated rats were studied. In control animals, [Dab9]N/OFQ(1–13)NH2 and [Dap9]N/OFQ(1–13)NH2, unlike N/OFQ(1–13)NH2 and [Cav9]N/OFQ(1–13)NH2, slightly increased the brain lipid peroxidation; the rest of the parameters were unchanged by all neuropeptides tested. KA (0.25 µg in 0.5 µl, i.c.v) increased the lipid peroxidation (4 and 24 h after KA‐injection) and decreased the glutathione level (1 h after KA‐administration). One hour after KA‐administration, the neuropeptides (2 µg in 0.5 µl, injected 30 min before KA) showed the following effects: a slight decrease in the KA‐induced lipid peroxidation by all nociceptin analogues and an enhancement of the KA‐decreased GSH level, but by [Cav9]N/OFQ(1–13)NH2 only. The brain antioxidant enzyme activities were unchanged in all used experimental groups. In addition, the nociceptin analogues, especially [Can9]N/OFQ(1–13)NH2, showed a good antioxidant capacity in chemical systems, generating reactive oxygen species. In conclusion, the substitution of lysin (Lys) in N/OFQ(1–13)NH2 molecule with other amino acids might contribute to changes in its antioxidant properties. Copyright © 2011 John Wiley & Sons, Ltd. 相似文献
5.
This study proposes a novel chemiluminescent assay of bacterial activity. Luminol chemiluminescence (LC) was amplified on addition of menadione to Escherichia coli suspension, and it was effectively inhibited by addition of superoxide dismutase rather than catalase. This fact suggests that H2O2 produced from O2 by superoxide dismutase is decomposed by catalase of E. coli. NAD(P)H:menadione reductase activities in periplasm and cytosol corresponded to the amplification of menadione-catalyzed LC, and outer and cytoplasmic membranes were only slightly involved in the LC. The total activity and Vmax of NAD(P)H:menadione reductase in the cytoplasm were greater than those in the periplasm. A transient increase in menadione-catalyzed LC was observed in the exponential phase and the LC decreased in the stationary phase during growth of E. coli. Menadione-catalyzed LC was sensitive to antibiotic action. A decrease in menadione-catalyzed LC by the impairment of membrane functions and by the inhibition of protein synthesis was observed at 5 min and 3 hr, respectively. These findings suggest the possibility that menadione-catalyzed luminol chemiluminescent assay is applicable to rapid antimicrobial assay because LC is sensitive to the change in growth and cytotoxic events caused by antimicrobial agents. 相似文献
6.
采用脱氨再生几丁质凝胶亲和层析从萝卜叶中获得了PAGE圆盘电泳均一的溶菌酶,并通过微量蒸发扩散法获得其结晶。用pH动力学研究方法和化学修饰法对该酶的活性中心进行了初步研究,实验结果表明:-COOH(Glu/Asp)基团、Trp及His残基可能为酶活性所必需。采用对Tyr、Cys、Arg、Ser/Thr残基专一的修饰剂对酶作用后,酶活性基本上不受影响。同时,酶活性受到组胺和GlcNAc抑制。讨论了萝卜溶菌酶与鸡蛋清溶菌酶和木瓜溶菌酶的活性中心的相互差异。 相似文献
7.
A quantitative structure-activity relationship (QSAR) study has been made on some series of anti-hepatitis B virus (HBV) agents, namely, a series of novel bis(L-amino acid) ester prodrugs of 9-[2--(phosphonomethoxy)ethyl]adenine, a similar series of compounds comprising of 2- amino-6-arylthio-9-[2-(phosphonoethoxy)ethyl] purine bis(2,2,2- trifluoroethyl) esters, and a series of 1-isopropylsulfonyl-2-amine benzimidazoles. In each case significant correlations are found between the anti-HBV potencies and some physicochemical and steric properties of the compounds, indicating that for the first two series the activity is controlled by the hydrophobic and the bulk properties of the molecules and, for the third series, the steric and hydrogen bonding properties of compounds are crucial for their anti-HBV potency. 相似文献
8.
Ahmed H. Moustafa Hassan A. El-Sayed Abd El-Fattah Z. Haikal El Sayed H. El Ashry 《Nucleosides, nucleotides & nucleic acids》2013,32(5):340-352
Reaction of pyridin-2(1H)-one 1 with 4-bromobutylacetate (2), (2-acetoxyethoxy)methyl bromide (3) gave the corresponding nicotinonitrile O-acyclonucleosides, 4 and 5, respectively. Deacetylation of 4 and 5 gave the corresponding deprotected acyclonucleosides 6 and 7, respectively. Treatment of pyridin-2(1H)-one 1 with 1,3-dichloropropan-2-ol (8), epichlorohydrin (10) and allyl bromide (12) gave the corresponding nicotinonitrile O-acyclonucleosides 9, 11, and 13, respectively. Furthermore, reaction of pyridin-2(1H)-one 1 with the propargyl bromide (14) gave the corresponding 2-O-propargyl derivative 15, which was reacted via [3+2] cycloaddition with 4-azidobutyl acetate (16) and [(2-acetoxyethoxy)methyl]azide (17) to give the corresponding 1,2,3-triazole derivatives 18 and 19, respectively. The structures of the new synthesized compounds were characterized by using IR, 1H, 13C NMR spectra, and microanalysis. Selected members of these compounds were screened for antibacterial activity. 相似文献
9.
Xi‐ou Xiao Wenqiu Lin Ke Li Wei Li Xiaomin Gao Lingling Lv 《Journal of Phytopathology》2017,165(10):652-661
Reactive oxygen species (ROS) play a crucial role in the early response to plant biotic and abiotic stresses. In this study, bacterial wilt‐resistant and wilt‐susceptible eggplants were inoculated with Ralstonia solanacearum and the ROS content was analysed. The result revealed an increased accumulation of hydrogen peroxide (H2O2) and superoxide (O2?) in resistant and susceptible eggplant roots after R. solanacearum inoculation. H2O2 and O2? accumulation increased earlier in the inoculated resistant eggplant root than in the inoculated susceptible eggplant root. Real‐time polymerase chain reaction results revealed that respiratory burst oxidase homologue (Rboh) A, RbohB, RbohF and PR1 expression levels increased in inoculated resistant eggplant roots at an early stage (0–60 h postinoculation) and were at higher expression levels than those in susceptible eggplant roots. Ascorbate peroxidase, peroxidase and catalase activities were higher in inoculated resistant eggplant roots than in susceptible eggplant roots at the early stage. Hence, an early ROS burst positively regulates bacterial wilt resistance in eggplant. 相似文献
10.
Although the development of cellular hypertrophy is widely believed to involve Ca(2+) signaling, potential supporting roles for sequestered Ca(2+) in this process have not been explored. H9c2 cardiomyocytes respond to arginine vasopressin with an initial mobilization of Ca(2+) stores and reduced rates of mRNA translation followed by repletion of Ca(2+) stores, up-regulation of translation beyond initial rates, and the development of hypertrophy. Rates of synthesis of the endoplasmic reticulum (ER) chaperones, GRP78 and GRP94, were found to increase preferentially at early times of vasopressin treatment. Total GRP78 content increased 2- to 3-fold within 8 h after which the chaperone was subject to post-translational modification. Preferential synthesis of GRP78 and the increase in chaperone content both occurred at pM vasopressin concentrations and were abolished at supraphysiologic Ca(2+) concentrations. Co-treatment with phorbol myristate acetate decreased vasopressin-dependent Ca(2+) mobilization and slowed appearance of new GRP78 molecules in response to the hormone, whereas 24 h pretreatment with phorbol ester prolonged vasopressin-dependent Ca(2+) mobilization and further increased rates of GRP78 synthesis in response to the hormone. Findings did not support a role for newly synthesized GRP78 in translational up-regulation by vasopressin. However up-regulation, which does not depend on Ca(2+) sequestration, appeared to expedite chaperone expression. This report provides the first evidence that a Ca(2+)-mobilizing hormone at physiologic concentrations signals increased expression of GRP78. Translational tolerance to depletion of ER Ca(2+) stores, typifying a robust ER stress response, did not accompany vasopressin-induced hypertrophy. 相似文献
11.
Ken Tsutsumi Katsunori Endou Akinori Furutani Tomomi Ikki Hiroaki Nakano Takuya Shintani Tsumoru Morimoto Kiyomi Kakiuchi 《Chirality》2003,15(6):504-509
The diastereoselective [2+2] photocycloaddition of cyclohexenonecarboxylates containing various chiral auxiliaries to ethylene is described. The effect of the auxiliary, reaction temperature, and solvent on diastereoselectivity was examined. The (?)‐8‐(p‐methoxyphenyl)menthyl group was found to be the most effective chiral auxiliary. The photoreaction of (?)‐8‐(p‐methoxyphenyl)menthyl cyclohexenonecarboxylate in methylcyclohexane at ?78°C gave the corresponding bicyclo[4.2.0]octanone derivative in 81% diastereomeric excess (d.e.). The extent of diastereoselectivity was found to be closely related to the most stable π‐stack conformation of the starting cyclohexenones. Chirality 15:504–509, 2003. © 2003 Wiley‐Liss, Inc. 相似文献
12.
Morgane Lardic Cedric Patry Muriel Duflos Jean Guillon Stephane Massip Francisco Cruzalegui 《Journal of enzyme inhibition and medicinal chemistry》2013,28(3):313-325
New series of 2(or 3)-arylmethylenenaphtho[2,1-b]furan-3(or 2)-ones were synthesized, characterized and tested for anticancer properties in vitro. The target compounds were prepared by Knoevenagel coupling between the naphthofuranones 3, 28–30 and formyl derivatives. 2-(4-Oxo-1-benzopyran-3-ylmethylene)naphtho[2,1-b]furan-3-one 36 was the most active compound (IC50 (L1210) = 1.6 μM). These compounds were also evaluated, in an independent manner, as inhibitors of Src protein tyrosine kinase, but only minor activity was observed. 相似文献
13.
Tritium-labeled hemicholinium-3 ([3H]HC-3) was used to characterize the sodium-dependent high-affinity choline carrier sites in rat striatal preparations. In an earlier study, we had shown that [3H]HC-3 labels choline carrier sites with high and low affinities and had suggested that the low-affinity sites represent "functional" carrier sites. The objective of the present study was to examine the mechanisms involved in the regulation of the two affinity states of [3H]HC-3 binding. Here, we demonstrate that these two affinity states are totally interconvertible; addition of 0.1 mM ATP in the binding assay medium quantitatively converted all the binding sites to the low-affinity state, whereas addition of 1 mM beta,gamma-methylene 5'-ATP quantitatively converted all the binding sites to the high-affinity state. Preincubation of the tissue (for 15 min at 37 degrees C) before the binding assay also converted the binding sites to the high-affinity state, whereas supplementation of the assay medium with ATP (0.5 mM) again induced expression of the low-affinity state of the binding sites. This effect of ATP was found to be selective for this nucleotide. Neither ADP (1 mM) nor cyclic AMP could mimic such an effect. Other nucleotide triphosphates--CTP (0.5 mM) and GTP (0.5 mM)--also could not substitute for ATP. GTP, however, caused nearly a 35% reduction in the number of binding sites, accompanying a loss of the low-affinity component of binding. This effect of GTP was also shared by 5'-guanylylimidodiphosphate but not by GDP or cyclic GMP. This ATP-dependent low-affinity conversion of [3H]HC-3 binding sites requires divalent metal ions.(ABSTRACT TRUNCATED AT 250 WORDS) 相似文献
14.
Borsuk K Grzeszczyk B Szczukiewicz P Przykorska B Frelek J Chmielewski M 《Chirality》2004,16(7):414-421
The [2+2]cycloaddition of CSI to the (Z)-propenyl ethers derived from respective 1,3-methylidene- and 1,3-ethylidene-threitols, contrary to the corresponding erythritol derivatives, is characterized by a low stereoselectivity and a lack of stereospecificity. On the other hand, the alternative method of the oxacepham formation, based on the 4-vinyloxy-azetidinone, proceeds with an excellent stereoselectivity. The CD-spectroscopy offers an attractive tool for determination of the absolute configuration of the bridgehead carbon atom at the 5-oxacepham skeleton. 相似文献
15.
《Luminescence》2002,17(2):117-122
The electrogenerated chemiluminescence of Ru(bpy)32+/C2O42? system on a pre‐polarized Au electrode was studied using a potential‐resolved electrochemiluminescence (PRECL) method. Two anodic ECL peaks were observed at 1.22 V (vs. SCE) (EP1), 1.41 V (vs. SCE) (EP2), respectively. The effects of the concentration of oxalate and Ru(bpy)32+, adsorbed sulphur, CO2, O2, pH of the solution and pretreatment of the Au electrode on the two PRECL peaks were examined. The surface state of the pre‐oxidized gold electrode was also studied using the X‐ray photoelectron spectroscopy (XPS) technique. Moreover, comparative studies on i–E and I–E curves were carried out and a possible mechanism involving both the catalytic and the direct electro‐oxidation pathways was proposed for the ECL of Ru(bpy)32+/C2O42? system. EP1 is attributed to the Ru(bpy)32/3+ reaction catalysed by C2O42? to generate Ru(bpy)32+*. EP2 is likely because C2O42? was oxidized at the electrode to form CO2?·, followed by reaction with Ru(bpy)33+ to generate Ru(bpy)32+*. Copyright © 2002 John Wiley & Sons, Ltd. 相似文献
16.
Elina Tzvetanova Almira Pavlova Albena Alexandrova Galina Nenkova Lubomir Petrov Margarita Kirkova Radoslav Girchev Emilia Naydenova 《Cell biochemistry and function》2009,27(4):243-250
In‐vivo effects of nociceptin (N/OFQ(1‐13)NH2) on the levels of lipid peroxidation and cell enzyme (superoxide dismutase, glutathione peroxidase and glutathione reductase) and non‐enzyme (glutathione) antioxidants in brain of control and kainic acid‐treated rats were studied. N/OFQ(1‐13)NH2 effects were compared with those of its structural analogue [Orn9]N/OFQ(1‐13)NH2. Kainic acid (25 µg, i.c.v) increased the lipid peroxidation (4 and 24 h after kainic acid treatment) and decreased the glutathione level (1 h after kainic acid injection). We failed to find, any changes in antioxidant enzyme activities, independently of the time of kainic acid treatment. At the background of kainic acid‐effects, N/OFQ(1‐13)NH2 and [Orn9] N/OFQ(1‐13)NH2, injected 30 min before kainic acid, had no effects on all parameters, tested in brain. In addition, the neuropeptides did not change the antioxidant status in brain of control animals. It might be concluded that N/OFQ(1‐13)NH2 and [Orn9]N/OFQ(1‐13)NH2 have neither pro‐ nor anti‐oxidant activity. Copyright © 2009 John Wiley & Sons, Ltd. 相似文献
17.
Platelet-derived growth factor (PDGF) is established to function importantly in the growth, development, and function of most cardiovascular tissues. However, evidence that the factor participates directly in the growth and development of the mammalian myocardium is lacking. H9c2 rat embryonic ventricular myocytes were found to respond to PDGF-BB with a rapid mobilization of cell-associated Ca2+ and increased rates of protein synthesis, followed by markedly increased rates of DNA synthesis. PDGF acted as a full mitogen for these myocytes. Evidence is provided that documents the expression of classical PDGF-beta, but not PDGF-alpha, receptors in H9c2 cells. Scatchard analysis revealed the presence of 44,000 beta-receptors per myocyte. Cell shortening and clustering of plasmalemmal beta-receptors occurred within 30 min of exposure to PDGF-BB. Treatment was also associated with a transient increase in the rate of synthesis of GRP78/BiP, consistent with a transitory release of Ca2+ from the sarcoplasmic/endoplasmic reticulum [S(E)R]. Increased rates of protein synthesis at early times of PDGF treatment were additive with those occurring in response to arginine vasopressin, indicating different mechanisms of translational upregulation by these agents. The mitogenic effects of PDGF were delayed by vasopressin, which causes H9c2 myocytes to undergo hypertrophy while promoting the persistent depletion of S(E)R Ca2+ stores. In the presence of PDGF, vasopressin did not induce hypertrophy. As compared to untreated myocytes, DNA synthesis in PDGF-treated myocytes was optimized at lower extracellular Ca2+ concentrations and was significantly less sensitive to inhibition by ionomycin. H9c2 cells appear to provide a useful embryonic cardiomyocyte model in which to examine both PDGF-activated proliferative and vasopressin-activated hypertrophic events and the importance of transient vs. sustained Ca2+ release in these events. 相似文献
18.
Chirally deuterated (S)-D-(6-(2)H(1))glucose has been prepared in good overall yield from d-(6,6'-(2)H(2))glucose by a short, five-step synthesis from D-(6,6-(2)H(2))glucose utilizing (R)-(+)-Alpine-Borane [(R)-9-[(6,6-dimethylbicyclo[3.1.1]hept-2-yl)methyl]-9-borabicyclo[3.3.1]nonane]. Suitably protected methyl 2,3,4-tri-O-benzyl-D-(6,6-(2)H(2))glucopyranoside was prepared and the deuterated O-6 primary alcohol was oxidized to an aldehyde by Swern oxidation. Stereoselective reduction with nondeuterated (R)-(+)-Alpine-Borane gave methyl 2,3,4-tri-O-benzyl-(6S)-D-(6-(2)H(1))glucopyranoside, which was deprotected under standard conditions to afford the title compound. The key stereoselective reduction step was achieved in 90% yield. The preparation uses economical, commercially available starting materials and will be useful for elucidating biosynthetic mechanisms. 相似文献
19.
A novel type of oxygenolytic ring cleavage: 2,4-oxygenation and decarbonylation of 1H-3-hydroxy-4-oxoquinaldine and 1H-3-hydroxy-4-oxoquinoline 总被引:1,自引:0,他引:1
Iris Bauer ré de Beyer Barbara Tshisuaka Susanne Fetzner Franz Lingens 《FEMS microbiology letters》1994,117(3):299-304
Abstract The utilization of quinaldine (2-methylquinoline) by Arthrobacter sp. Rü61a proceeds via 1 H -4-oxoquinaldine, 1 H -3-hydroxy-4-oxoquinaldine, and N -acetyl-anthranilic acid. By analogy, 1 H -4-oxoquinoline is degraded by Pseudomonas putida 33/1 via 1 H -3-hydroxy-4-oxoquinoline and N -formylanthranilic acid. Using the purified enzymes from both organisms, the mode of N -heterocyclic ring cleavage was investigated. The conversions of 1 H -3-hydroxy-4-oxoquinaldine and 1 H -3-hydroxy-4-oxoquinoline to N -acetyl- and N -formylanthranilic acid, respectively, were both accompanied by the release of carbon monoxide. The enzyme-catalysed transformations were performed in an [18 O]O2 atmosphere and resulted in the incorporation of two oxygen atoms into the respective products, N -acetyl- and N -formylanthranilic acid, indicating an oxygenolytic attack at C-2 and C-4 of both 1 H -3-hydroxy-4-oxoquinaldine and 1 H -3-hydroxy-4-oxoquinolone. 相似文献