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 0 毫秒
1.
《Free radical research》2013,47(1-2):63-68
α-2.6-Difluorophenyl-N-tert-butylnitrone (F2PBN) was synthesized and evaluated. A number of alkyl adducts of F2PBN were studied by ESR/ENDOR. An additional hyperfine splitting (a triplet of doublets of doublets) is reported. The existence of two (one large, one small) F-hfsc's from the ortho-fluorine atoms in the phenyl ring of most alkyl adducts was confirmed by ENDOR spectroscopy.  相似文献   

2.
LiMgBO3:Dy3+, a low Zeff material was prepared using the solution combustion method and its luminescence properties were studied using X‐ray diffraction (XRD), scanning electron microscopy (SEM), thermoluminescence (TL), photoluminescence (PL), Fourier transform infrared spectroscopy, and electron paramagnetic resonance (EPR) techniques. Reitvield refinement was also performed for the structural studies. The PL emission spectra for LiMgBO3:Dy3+ consisted of two peaks at 478 due to the 4F9/26H15/2 magnetic dipole transition and at 572 nm due to the hypersensitive 4F9/26H13/2 electric dipole transition of Dy3+, respectively. A TL study was carried out for both the γ‐ray‐irradiated sample and the C5+ irradiated samples and was found to show high sensitivity for both. Moreover the γ‐ray‐irradiated LiMgBO3:Dy3+ sample showed linearity in the dose range 10 Gy to 1 kGy and C5+‐irradiated samples show linearity in the fluence range 2 × 1010 to 1 × 1011 ions/cm2. In the present study, the initial rise method, various heating rate method, the whole glow curve method, glow curve convolution deconvolution function, and Chen's peak shape method were used to calculate kinetic parameters to understand the TL glow curve mechanism in detail. Finally, an EPR study was performed to examine the radicals responsible for the TL process.  相似文献   

3.
Immunoglobin G with α‐2,6 sialylation has been reported to have an impact on antibody‐dependent cellular cytotoxicity and anti‐inflammatory efficacy. However, production of antibodies with α‐2,6 sialylation from Chinese hamster ovary cells is challenging due to the inaccessibility of sialyltransferases for the heavy chain N‐glycan site and the presence of exclusively α‐2,3 sialyltransferases. In this study, combining mutations on the Fc regions to allow sialyltransferase accessibility with overexpression of α‐2,6 sialyltransferase produced IgG with significant levels of both α‐2,6 and α‐2,3 sialylation. Therefore, ST3GAL4 and ST3GAL6 genes were disrupted by CRISPR/Cas9 to minimize the α‐2,3 sialylation. Sialidase treatment and SNA lectin blot indicated greatly increased α‐2,6 sialylation level relative to α‐2,3 sialylation for the α‐2,3 sialyltransferase knockouts when combined with α‐2,6 sialyltransferase overexpression. Indeed, α‐2,3 linked sialic acids were not detected on IgG produced from the α‐2,3 sialyltransferase knockout‐α‐2,6 sialyltransferase overexpression pools. Finally, glycoprofiling of IgG with four amino acid substitutions expressed from an α‐2,3 sialyltransferase knockout‐α‐2,6 sialyltransferase stable clone resulted in more than 77% sialylated glycans and more than 62% biantennary disialylated glycans as indicated by both MALDI‐TOF and LC‐ESI‐MS. Engineered antibodies from these modified Chinese hamster ovary cell lines will provide biotechnologists with IgGs containing N‐glycans with different structural variations for examining the role of glycosylation on protein performance.  相似文献   

4.
A semi-micro assay was developed for the conjugation of 5α,6α-epoxy-cholestan-3β-ol (cholesterol α-oxide) with glutathione. The soluble supernatant of rat liver homogenate catalysed the reaction at a rate of 0.2–0.5 pmol.min−1 .mg protein−1 with 4μM cholesterol α-oxide, while the reaction in the presence of GSH alone was barely detectable. Enzymic activity in the soluble supernatant was due equally to the two forms of glutathione transferase B (100 pmol.min.mg protein−1), glutathione transferases AA, A, C and E being unreactive. The activity of purified glutathione transferase B was about 5-times that expected from the activity of the soluble supernatant. Complex enzyme kinetics were obtained suggestive of substrate inhibition.  相似文献   

5.
6.
β1,4 galactosyl- and α2,6 sialyltransferase (gal-T EC 2.4.1.22 and sialyl-T EC 2.4.99.1) sequentially elongate and terminate complex N-glycan chains of glycoproteins. Both enzymes reside in trans Golgi cisternae; their ultrastructural relationship, however, is unknown. To delineate their respective Golgi compartment(s) we conducted a double label immunofluorescent study by conventional and confocal laser scanning microscopy in HepG2, HeLa, and other cells in presence of Golgi-disturbing agents. Polyclonal, peptide-specific antibodies to human sialyl-T expressed as a β-galactosidase-sialyl-T fusion protein in E. coli were developed and applied together with mABs to human milk gal-T. In untreated HepG2 and HeLa cells Golgi morphology identified by immunofluorescent labeling of sialyl-T and gal-T, respectively, was nearly identical. Treatment of cells with brefeldin A (BFA) led to rapid and coordinated disappearance of immunostaining of both enzymes; after BFA washout, vesicular structures reappeared which first stained for gal-T followed by sialyl-T; in the reassembled Golgi apparatus sialyl-T and gal-T were co-localized again. In contrast, monensin treatment produced a reversible swelling and scattering of gal-T positive Golgi elements while sialyl-T positive structures showed little change. Treatment with nocodazole led to dispersal of Golgi elements in which gal-T and sialyl-T remained co-localized. Treatment with chloroquine affected Golgi structure less than monensin and led to condensation of gal-T positive and to slight enlargement of sialyl-T positive structures. Sequential recovery from BFA of gal-T and sialyl-T and their segregation by monensin suggest that these enzymes are targeted to different Golgi subcompartments.  相似文献   

7.
Interactions of cell-adhesion molecule LFA-1 and its ligand ICAM-1 play important roles during immune and inflammatory responses. Critical residues of LFA-1 for ICAM-1 binding are known to be in the I-domain of the α-subunit and the I-like domain of the β-subunit. On the basis of our previous work demonstrating the inhibitory activity of I-domain cyclic peptide cLAB.L on LFA-1/ICAM-1 interaction, here we have explored the activity of I-like-domain peptide LBE on the binding mechanism of cLAB.L. LBE enhances cLAB.L binding to T-cells and epithelial cells. The adherence of T-cells to epithelial monolayers was suppressed by the two peptides. The addition of LBE to the monolayers prior to the addition cLAB.L produced a better inhibitory effect than the reverse procedure. LBE, but not cLAB.L, changes the ICAM-1 conformation, suggesting that LBE binds to ICAM-1 at sites that are distinct from these of cLAB.L and induces improved conformation in ICAM-1 for binding to cLAB.L.  相似文献   

8.
The actions of the hormonal form of vitamin D, 1α,25-dihydroxyvitamin D3 [1α,25-(OH)2D3], are mediated by both genomic and nongenomic mechanisms. Several vitamin D synthetic analogs have been developed in order to identify and characterize the site(s) of action of 1α,25-(OH)2D3 in many cell types including osteoblastic cells. We have compared the effects of 1α,25-(OH)2D3 and a novel 1α,25-(OH)2D3 bromoester analog (1,25-(OH)2-BE) that covalently binds to vitamin D receptors. Rat osteosarcoma cells that possess (ROS 17/2.8) or lack (ROS 24/1) the classic intracellular vitamin D receptor were studied to investigate genomic and nongenomic actions. In ROS 17/2.8 cells plated at low density, the two vitamin D compounds (1 × 10−8 M) caused increased cell proliferation, as assessed by DNA synthesis and total cell counts. Northern blot analysis revealed that the mitogenic effect of both agents was accompanied by an increase in steady-state osteocalcin mRNA levels, but neither agent altered alkaline phosphatase mRNA levels in ROS 17/2.8 cells. ROS 17/2.8 cells responded to 1,25-(OH)2-BE but not the natural ligand with a significant increase in osteocalcin secretion after 72, 96, 120, and 144 hr of treatment. Treatment of ROS 17/2.8 cells with the bromoester analog also resulted in a significant decrease in alkaline phosphatase-specific activity. To compare the nongenomic effects of 1α,25-(OH)2D3 and 1,25-(OH)2-BE, intracellular calcium was measured in ROS 24/1 cells loaded with the fluorescent calcium indicator Quin 2. At 2 × 10−8 M, both 1α,25-(OH)2D3 and 1,25-(OH)2-BE increased intracellular calcium within 5 min. Both the genomic and nongenomic actions of 1,25-(OH)2-BE are similar to those of 1α,25-(OH)2D3, and since 1,25-(OH)2-BE has more potent effects on osteoblast function than the naturally occurring ligand due to more stable binding, this novel vitamin D analog may be useful in elucidating the structure and function of cellular vitamin D receptors. © 1996 Wiley-Liss, Inc.  相似文献   

9.
The present work describes three novel nonpolar host peptide sequences that provide a ready assessment of the 310- and α-helix compatibilities of natural and unnatural amino acids at different positions of small- to medium-size peptides. The unpolar peptides containing Ala, Aib, and a C-terminal p-iodoanilide group were designed in such a way that the peptides could be rapidly assembled in a modular fashion, were highly soluble in solvent mixtures of triflouroethanol and H2O for CD- and two-dimensional (2D) nmr spectroscopic analyses, and showed excellent crystallinity suited for x-ray structure analysis. To validate our approach we synthesized 9-mer peptides 79a–96 (Table IV), 12-mer peptides 99–110c (Table V), and 10-mer peptides 120a–125d and 129–133 (Table VI and Scheme 8) incorporating a series of optically pure cyclic and open-chain (R)- and (S)-α,α-disubstituted glycines 1–10 (Figure 2). These amino acids are known to significantly modulate the conformations of small peptides. Based on x-ray structures of 9-mers 79a, 80, and 87 (Figures 4–7), 10-mers 124c, 131, and 132 (Figures 9–12), and 12-mer peptide 102b (Figure 13), CD spectra of all peptides recorded in acidic, neutral, and basic media and detailed 2D-nmr analyses of 9-mer peptide 86 and 12-mer 102b, several interesting conformational observations were made. Especially interesting results were obtained using the convex constraint CD analysis proposed by Fasman on 9-mer peptides 79a–d, 80, 81, 86, and 87, which allowed us to determine the relative content of 310- and α-helical conformations. These results were fully supported by the corresponding x-ray and 2D-nmr analyses. As a striking example we found that the (S)- and (R)-β-tetralin derived amino acids (R)- and (S)-1 show excellent α-helix stabilisation, more pronounced than Aib and Ala. These novel reference peptide sequences should help establish a scale for natural and unnatural amino acids concerning their intrinsic 310- and α-helix compatibilities at different positions of medium-sized peptides and thus improve our understanding in the folding processes of peptides. © 1997 John Wiley & Sons, Inc. Biopoly 42: 575–626, 1997  相似文献   

10.
It has been suggested that part of the increased β-catecholamine responsiveness in hyperthyroid animals is due to a decrease in α-catecholamine action. The present results indicate that neither hyperthyroidism nor hypothyroidism altered the α2-adrenergic inhibition of adenylate cyclase or the α1-adrenergic stimulation of phosphatidylinositol turnover in adipocytes from the white adipose tissue of hamster. No effect of hyperthyroidism was found on the Kd of [3H]dihydroegocryptine or the number of binding sites in membranes prepared from hamster adipocyte tissue. The stimulation of cyclic AMP due to β-catecholamines was enhanced in adipocytes from hyperthyroid hamster, as was lipolysis. However, in adipocytes from hyperthyroid hamster the maximal stimulation of cyclic AMP due to isoproterenol, ACTH or epinephrine plus yohimbine, as seen in the presence of adenosine deaminase and theophylline, was less than in adipocytes from euthyroid hamsters. The activation of adenylate cyclase by isoproterenol was the same in membranes from hyperthyroid as compared to those from euthyroid hamsters in the absence or presence of guanine nucleotides. These data suggest that thyroid status has little effect on α-catecholamine action but enhances the activation of lipolysis by β-catecholamine agonists.  相似文献   

11.
Herein, the authors present the development of novel 0D–2D nanohybrids consisting of a nickel‐based bimetal phosphorus trisulfide (Ni1?xFexPS3) nanomosaic that decorates on the surface of MXene nanosheets (denoted as NFPS@MXene). The nanohybrids are obtained through a facile self‐assemble process of transition metal layered double hydroxide (TMLDH) on MXene surface; followed by a low temperature in situ solid‐state reaction step. By tuning the Ni:Fe ratio, the as‐synthesized NFPS@MXene nanohybrids exhibit excellent activities when tested as electrocatalysts for overall water splitting. Particularly, with the initial Ni:Fe ratio of 7:3, the obtained Ni0.7Fe0.3PS3@MXene nanohybrid reveals low overpotential (282 mV) and Tafel slope (36.5 mV dec?1) for oxygen evolution reaction (OER) in 1 m KOH solution. Meanwhile, the Ni0.9Fe0.1PS3@MXene shows low overpotential (196 mV) for the hydrogen evolution reaction (HER) in 1 m KOH solution. When integrated for overall water splitting, the Ni0.7Fe0.3PS3@MXene || Ni0.9Fe0.1PS3@MXene couple shows a low onset potential of 1.42 V and needs only 1.65 V to reach a current density of 10 mA cm?2, which is better than the all noble metal IrO2 || Pt/C electrocatalyst (1.71 mV@10 mA cm?2). Given the chemical versatility of Ni1?xFexPS3 and the convenient self‐assemble process, the nanohybrids demonstrated in this work are promising for energy conversion applications.  相似文献   

12.
Dehydroamino acids are non‐coded amino acids that offer unique conformational properties. Dehydrophenylalanine (ΔPhe) is most commonly used to modify bioactive peptides to constrain the topography of the phenyl ring in the side chain, which commonly serves as a pharmacophore. The Ramachandran maps (in the gas phase and in CHCl3 mimicking environments) of ΔPhe analogues with methyl groups at the β position of the side chain as well as at the C‐terminal amide were calculated using the B3LYP/6‐31 + G** method. Unexpectedly, β‐methylation alone results in an increase of conformational freedom of the affected ΔPhe residue. However, further modification by introducing an additional methyl group at C‐terminal methyl amide results in a steric crowding that fixes the torsion angle ψ of all conformers to the value 123°, regardless of the Z or E position of the phenyl ring. The number of conformers is reduced and the accessible conformational space of the residues is very limited. In particular, (Z)‐Δ(βMe)Phe with the tertiary C‐terminal amide can be classified as the amino acid derivative that has a single conformational state as it seems to adopt only the β conformation. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
A good model to experimentally explore evolutionary hypothesis related to enzyme function is the ancient‐like dual‐substrate (βα)8 phosphoribosyl isomerase A (PriA), which takes part in both histidine and tryptophan biosynthesis in Streptomyces coelicolor and related organisms. In this study, we determined the Michaelis–Menten enzyme kinetics for both isomerase activities in wild‐type PriA from S. coelicolor and in selected single‐residue monofunctional mutants, identified after Escherichia coli in vivo complementation experiments. Structural and functional analyses of a hitherto unnoticed residue contained on the functionally important β → α loop 5, namely, Arg139, which was postulated on structural grounds to be important for the dual‐substrate specificity of PriA, is presented for the first time. Indeed, enzyme kinetics analyses done on the mutant variants PriA_Ser81Thr and PriA_Arg139Asn showed that these residues, which are contained on β → α loops and in close proximity to the N‐terminal phosphate‐binding site, are essential solely for the phosphoribosyl anthranilate isomerase activity of PriA. Moreover, analysis of the X‐ray crystallographic structure of PriA_Arg139Asn elucidated at 1.95 Å herein strongly implicates the occurrence of conformational changes in this β → α loop as a major structural feature related to the evolution of the dual‐substrate specificity of PriA. It is suggested that PriA has evolved by tuning a fine energetic balance that allows the sufficient degree of structural flexibility needed for accommodating two topologically dissimilar substrates—within a bifunctional and thus highly constrained active site—without compromising its structural stability.  相似文献   

14.
Oxidation of the reactive site methionine (Met) in α-1-proteinase inhibitor (α-1-PI) to methionine sulfoxide (Met(O)) is known to cause depletion of its elastase inhibitory activity. To estimate the selectivity of different oxidants in converting Met to Met(O) in α-1-PI, we measured the molar ratio Met(O)/α-1-PI at total inactivation. This ratio was determined to be 1.2 for both the myeloperoxidase/H2O2/chloride system and the related compound NH2Cl. With taurine monochloramine, another myeloperoxidase-related oxidant, 1.05 mol Met(O) were generated per mol α-1-PI during inactivation. These oxidants attack preferentially one Met residue in α-1-PI, which is identical with Met 358, as concluded from the parallelism of loss of elastase inhibitory activity and oxidation of Met. A similar high specificity for Met oxidation was determined for the xanthine oxidase-derived oxidants. In contrast, the ratio found for ozone and m-chloroperoxybenzoic acid was 6.0 and 5.0, respectively, indicating oxidation of additional Met residues besides the reactive site Met in α-1-PI, i.e. unselective action of these oxidants. Further studies were performed on the efficiency of oxidants for total depletion of the elastase inhibitory capacity of α-1-PI. Ozone and m-chloroperoxybenzoic acid were 10-fold less effective and the superoxide anion/hydroxyl radicals were 30–50-fold less effective to inactivate the elastase inhibitory activity as compared to the myeloperoxidase-derived oxidants. The myeloperoxidase-related oxidants are discussed as important regulators of α-1-PI activity in vivo.  相似文献   

15.
The effect of α- and β-tocopherol on human erythroleukemia cell (HEL) adhesion induced by phorbol 12-myristate 13-acetate (PMA) has been studied. Adhesion induced by PMA stimulation was prevented by 44.5% by physiological concentrations of α-tocopherol. Under the same experimental conditions, β-tocopherol, an analogue of α-tocopherol, produced 11% inhibition of adhesion. Cell response gradually increased from 0 to 24 h of α-tocopherol treatment. Only a slight time dependency of β-tocopherol inhibition was observed. Another human erythroleukemia cell line (K562) and the human monocyte tumor cell line U937 showed 5.0 and 11.2% inhibition, respectively. Similar to α-tocopherol, the protein kinase C inhibitor, Calphostin C, and the MAPK inhibitor, PD98059, prevented PMA-induced cell adhesion. An inhibition of ERK-1 phosphorylation was observed for α-tocopherol only in HEL, implying that MAP kinase pathway is involved in this cell line. Fluorescence-activated cell sorting (FACS), by using various integrin-specific monoclonal antibodies, has shown that α (1–6), β1, and αv integrins are less expressed at the cell surface after α-tocopherol treatment. Beta-tocopherol treatment was less effective.  相似文献   

16.
17.
Introduction – Bioautographic assays using TLC play an important role in the search for active compounds from plants. A TLC assay has previously been established for the detection of β‐glucosidase inhibitors but not for α‐glucosidase. Nonetheless, α‐glucosidase inhibition is an important target for therapeutic agents against of type 2 diabetes and anti‐viral infections. Objective – To develop a TLC bioautographic method to detect α‐ and β‐glucosidase inhibitors in plant extracts. Methodology – The enzymes α‐ and β‐d ‐glucosidase were dissolved in sodium acetate buffer. After migration of the samples, the TLC plate was sprayed with enzyme solution and incubated at room temperature for 60 min in the case of α‐d ‐glucosidase, and 37°C for 20 min in the case of β‐d ‐glucosidase. For detection of the active enzyme, solutions of 2‐naphthyl‐α‐D‐glucopyranoside or 2‐naphthyl‐β‐D‐glucopyranoside and Fast Blue Salt were mixed at a ratio of 1 : 1 (for α‐d ‐glucosidase) or 1 : 4 (for β‐d ‐glucosidase) and sprayed onto the plate to give a purple background colouration after 2–5 min. Results – Enzyme inhibitors were visualised as white spots on the TLC plates. Conduritol B epoxide inhibited α‐d ‐glucosidase and β‐d ‐glucosidase down to 0.1 µg. Methanol extracts of Tussilago farfara and Urtica dioica after migration on TLC gave enzymatic inhibition when applied in amounts of 100 µg for α‐glucosidase and 50 µg for β‐glucosidase. Conclusion – The screening test was able to detect inhibition of α‐ and β‐glucosidases by pure reference substances and by compounds present in complex matrices, such as plant extracts. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
A sensitive fluorogenic probe 1 for biothiols was developed based on the Michael addition reaction. The probe 1 was readily synthesized via the reaction of 2‐(4′‐hydroxyphenyl) benzimidazole (HPBI) with acryloyl chloride and shows weak fluorescence emission. Upon mixing with biothiols, the fluorescence of 1 is significantly enhanced due to the conjugate addition of thiols to the α,β‐unsaturated carbonyl moiety, thus eliminating the photoinduced electron transfer (PET) quenching of the fluorophore by the intramolecular carbon–carbon double bond. Cysteine (Cys) was selected as the representative thiol in the spectral experiment. A good linear relationship was obtained from 1.0 to 30.0 µmol L?1 for Cys and the detection limit was 0.17 µmol L?1. Furthermore, probe 1 was highly selective for biothiols without the interference of some biologically relevant analytes and has been applied to detecting biothiols in human urine samples. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

19.
20.
Hematite (α‐Fe2O3) is widely used as a catalytic electrode material in photo‐electrochemical water oxidation, where its surface compositions and stabilities can strongly impact the redox reaction process. Here, its surface configurations in environmental or electrochemical conditions are assessed via density functional theory (DFT) calculations conducted at the Perdew, Burke, and Ernzerhof (PBE)+U level. The most energetically favorable surface domains of α‐Fe2O3 (0001) and (102) are predicted by constructing the surface phase diagrams in the framework of first‐principle thermodynamics. The relative surface stabilities are investigated as a function of partial pressures of oxygen and water, temperature, solution pH, and electrode potential not only for perfect bulk terminations but also for defect‐containing surfaces having various degrees of hydroxylation and hydration. In order to assess the impact on the redox reactions of the surface planes as well as of the extent of surface hydration/hydroxylation, the thermodynamics of the four‐step oxygen evolution reaction (OER) mechanism are examined in detail for different models of the α‐Fe2O3 (0001) and (102) surfaces. Importantly, the results underline that the nature of the surface termination and the degree of near‐surface hydroxylation give rise to significant variations in the OER overpotentials.  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