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.
Adriamycin, an anthracycline antibiotic, has been used for the treatment of various types of tumours. Adriamycin induces at least two distinct types of growth repression, such as senescence and apoptosis, in a concentration‐dependent manner. Cellular senescence is a condition in which cells are unable to proliferate further, and senescent cells frequently show polyploidy. Although abrogation of cell division is thought to correlate with polyploidization, the mechanisms underlying induction of polyploidization in senescent cells are largely unclear. We wished, therefore, to explore the role of cyclin B1 level in polyploidization of Adriamycin‐induced senescent cells. A subcytotoxic concentration of Adriamycin induced polyploid cells having the features of senescence, such as flattened and enlarged cell shape and activated β‐galactosidase activity. In DNA damage‐induced senescent cells, the levels of cyclin B1 were transiently increased and subsequently decreased. The decrease in cyclin B1 levels occurred in G2 cells during polyploidization upon treatment with a subcytotoxic concentration of Adriamycin. In contrast, neither polyploidy nor a decrease in cyclin B1 levels was induced by treatment with a cytotoxic concentration of Adriamycin. These results suggest that a decrease in cyclin B1 levels is induced by DNA damage, resulting in polyploidization in DNA damage‐induced senescence.  相似文献   

2.
Introduction – Flavonoids, the primary constituents of the petals of Nelumbo nucifera, are known to have antioxidant properties and antibacterial bioactivities. However, efficient methods for the preparative isolation and purification of flavonoids from this plant are not currently available. Objective – To develop an efficient method for the preparative isolation and purification of flavonoids from the petals of N. nucifera by high‐speed counter‐current chromatography (HSCCC). Methodology – Following an initial clean‐up step on a polyamide column, HSCCC was utilised to separate and purify flavonoids. Purities and identities of the isolated compounds were established by HPLC‐PAD, ESI‐MS, 1H‐NMR and 13C‐NMR. Results – The separation was performed using a two‐phase solvent system composed of ethyl acetate–methanol–water–acetic acid (4 : 1 : 5 : 0.1, by volume), in which the upper phase was used as the stationary phase and the lower phase was used as the mobile phase at a flow‐rate of 1.0 mL/min in the head‐to‐tail elution mode. Ultimately, 5.0 mg syringetin‐3‐O‐β‐d‐glucoside, 6.5 mg quercetin‐3‐O‐β‐d‐glucoside, 12.8 mg isorhamnetin‐3‐O‐β‐d‐glucoside and 32.5 mg kaempferol‐3‐O‐β‐d‐glucoside were obtained from 125 mg crude sample. Conclusion – The combination of HSCCC with a polyamide column is an efficient method for the preparative separation and purification of flavonoids from the petals of N. nucifera. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

3.
High‐throughput screens that dispense with the need for expensive synthetic Aβ peptide would be invaluable for identifying novel anti‐aggregants as potential treatments for Alzheimer's disease. A biosynthetic in vivo approach, using a recombinant fluorescent green fluorescent protein (GFP) reporter for the aggregation state of Aβ in Escherichia coli, has been reported by other workers. Here, inducible Aβ–GFP expression in E. coli was coupled to the concurrent constitutive production of a quasi‐random peptide library to screen for anti‐aggregant activity. To attempt to introduce greater robustness, mCherry was also co‐expressed as an internal fluorescence standard to allow ratiometric comparison between samples. However, fluctuations in mCherry expression levels, as well as a low dynamic range of GFP output between positive and negative anti‐aggregant peptides, highlighted limitations with the approach. Despite this, two novel peptides were identified that showed an equivalent in vitro anti‐aggregant activity to that of epigallocatechin‐3‐gallate. Thus, although biosynthetic in vivo strategies show promise as screens for novel activities, unforeseen problems can arise because of the variability inherent in any biological system. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
The conformation of oligomers of β‐amino acids of the general type Ac‐[β‐Xaa]n‐NHMe (β‐Xaa = β‐Ala, β‐Aib, and β‐Abu; n = 1–4) was systematically examined at different levels of ab initio molecular orbital theory (HF/6‐31G*, HF/3‐21G). The solvent influence was considered employing two quantum‐mechanical self‐consistent reaction field models. The results show a wide variety of possibilities for the formation of characteristic elements of secondary structure in β‐peptides. Most of them can be derived from the monomer units of blocked β‐peptides with n = 1. The stability and geometries of the β‐peptide structures are considerably influenced by the side‐chain positions, by the configurations at the Cα‐ and Cβ‐atoms of the β‐amino acid constituents, and especially by environmental effects. Structure peculiarities of β‐peptides, in particular those of various helix alternatives, are discussed in relation to typical elements of secondary structure in α‐peptides. © 1999 John Wiley & Sons, Inc. Biopoly 50: 167–184, 1999  相似文献   

5.
6.
7.
8.
The active site of ß‐galactosidase (E. coli) contains a Mg2+ ion ligated by Glu‐416, His‐418 and Glu‐461 plus three water molecules. A Na+ ion binds nearby. To better understand the role of the active site Mg2+ and its ligands, His‐418 was substituted with Asn, Glu and Phe. The Asn‐418 and Glu‐418 variants could be crystallized and the structures were shown to be very similar to native enzyme. The Glu‐418 variant showed increased mobility of some residues in the active site, which explains why the substitutions at the Mg2+ site also reduce Na+ binding affinity. The Phe variant had reduced stability, bound Mg2+ weakly and could not be crystallized. All three variants have low catalytic activity due to large decreases in the degalactosylation rate. Large decreases in substrate binding affinity were also observed but transition state analogs bound as well or better than to native. The results indicate that His‐418, together with the Mg2+, modulate the central role of Glu‐461 in binding and as a general acid/base catalyst in the overall catalytic mechanism. Glucose binding as an acceptor was also dramatically decreased, indicating that His‐418 is very important for the formation of allolactose (the natural inducer of the lac operon).  相似文献   

9.
Acetate formation is a disadvantage in the use of Escherichia coli for recombinant protein production, and many studies have focused on optimizing fermentation processes or altering metabolism to eliminate acetate accumulation. In this study, E. coli MEC697 (MG1655 nadR nudC mazG) maintained a larger pool of NAD(H) compared to the wild‐type control, and also accumulated lower concentrations of acetate when grown in batch culture on glucose. In steady‐state cultures, the elevated total NAD(H) found in MEC697 delayed the threshold dilution rate for acetate formation to a growth rate of 0.27 h?1. Batch and fed‐batch processes using MEC697 were examined for the production of β‐galactosidase as a model recombinant protein. Fed‐batch culture of MEC697/pTrc99A‐lacZ compared to MG1655/pTrc99A‐lacZ at a growth rate of 0.22 h?1 showed only a modest increase of protein formation. However, 1 L batch growth of MEC697/pTrc99A‐lacZ resulted in 50% lower acetate formation compared to MG1655/pTrc99A‐lacZ and a two‐fold increase in recombinant protein production.  相似文献   

10.
The increasing interest in click chemistry and its use to stabilize turn structures led us to compare the propensity for β‐turn stabilization of different analogs designed as mimics of the β‐turn structure found in tendamistat. The β‐turn conformation of linear β‐amino acid‐containing peptides and triazole‐cyclized analogs were compared to ‘conventional’ lactam‐ and disulfide‐bridged hexapeptide analogs. Their 3D structures and their propensity to fold in β‐turns in solution, and for those not structured in solution in the presence of α‐amylase, were analyzed by NMR spectroscopy and by restrained molecular dynamics with energy minimization. The linear tetrapeptide Ac‐Ser‐Trp‐Arg‐Tyr‐NH2 and both the amide bond‐cyclized, c[Pro‐Ser‐Trp‐Arg‐Tyr‐D ‐Ala] and the disulfide‐bridged, Ac‐c[Cys‐Ser‐Trp‐Arg‐Tyr‐Cys]‐NH2 hexapeptides adopt dominantly in solution a β‐turn conformation closely related to the one observed in tendamistat. On the contrary, the β‐amino acid‐containing peptides such as Ac‐(R)‐β3‐hSer‐(S)‐Trp‐(S)‐β3‐hArg‐(S)‐β3‐hTyr‐NH2, and the triazole cyclic peptide, c[Lys‐Ser‐Trp‐Arg‐Tyr‐βtA]‐NH2, both specifically designed to mimic this β‐turn, do not adopt stable structures in solution and do not show any characteristics of β‐turn conformation. However, these unstructured peptides specifically interact in the active site of α‐amylase, as shown by TrNOESY and saturation transfer difference NMR experiments performed in the presence of the enzyme, and are displaced by acarbose, a specific α‐amylase inhibitor. Thus, in contrast to amide‐cyclized or disulfide‐bridged hexapeptides, β‐amino acid‐containing peptides and click‐cyclized peptides may not be regarded as β‐turn stabilizers, but can be considered as potential β‐turn inducers. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
This work presents a comparative study of proton transfer activity (PTA) of two cationic (+6) antimicrobial peptides, β‐structural arenicin‐2 and α‐helical melittin. A new approach was proposed for the detection of passive proton transfer by using proteoliposomes containing bacteriorhodopsin, which creates a small light‐induced electrochemical proton gradient ?ΔpH. Addition of several nanomoles of the peptides lowers ?ΔpH that is proximately indicative of the pore formation. The quantitative analysis of sigmoidal dependences of ?pH on the peptides concentration was carried out using liposomes prepared from PC, PC/PE, PC/PE/PI and PC/PG. Substitution of PC‐containing liposomes with PE‐containing ones, having negative spontaneous curvature, reduced the PTA of α‐helical melittin and increased that of β‐structural arenicin‐2. This result indicates an essential difference in the pore formation by these peptides. Further increase of PTA in response to arenicin‐2 (in contrast to melittin) was observed in the liposomes prepared from PC/PE/PI. The data analysis leads to the conclusion that PTA is influenced by (i) efficiency of the pore assemblage, which depends on the structure of pore‐forming peptides, and the spontaneous curvature of lipids and (ii) the presence of mobile protons in the polar head groups of phospholipids. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
XPF‐St7 (GLLSNVAGLLKQFAKGGVNAVLNPK) is an antimicrobial peptide isolated from Silurana tropicalis. We developed an α‐helical segment of XPF‐St7 termed as XPF2. Using the XPF2 as a framework, we increased the positive net charge of XPF2 by amino acid substitutions, and thus obtained two novel antimicrobial peptides XPF4 and XPF6. These were each fused with an ubiquitin tag and successfully expressed in Escherichia coli. This ubiquitin fusion system may present a viable alternative for industrial production of antimicrobial peptides. XPF4 and XPF6 showed much better overall antimicrobial activity against both Gram‐negative and Gram‐positive bacteria than XPF2. The therapeutic index of XPF4 and XPF6 was 5.6‐fold and 6.7‐fold of XPF2, respectively. Bacterial cell membrane permeabilization and genomic DNA interaction assays were utilized to explore the mechanism of action of XPF serial peptides. The results revealed that the target of these antimicrobial peptides was the bacterial cytoplasmic membrane. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

13.
Three linear peptides incorporating d ‐Phe‐2‐Abz as the turn motif are reported. Peptide 1 , a hydrophobic β‐hairpin, served as a proof of principle for the design strategy with both NMR and CD spectra strongly suggesting a β‐hairpin conformation. Peptides 2 and 3, designed as amphipathic antimicrobials, exhibited broad spectrum antimicrobial activity, with potency in the nanomolar range against Staphylococcus aureus. Both compounds possess a high degree of selectivity, proving non‐haemolytic at concentrations 500 to 800 times higher than their respective minimal inhibitory concentrations (MICs) against S. aureus. Peptide 2 induced cell membrane and cell wall disintegration in both S. aureus and Pseudomonas aeruginosa as observed by transmission electron microscopy. Peptide 2 also demonstrated moderate antifungal activity against Candida albicans with an MIC of 50 μM. Synergism was observed with sub‐MIC levels of amphotericin B (AmB), leading to nanomolar MICs against C. albicans for peptide 2 . Based on circular dichroism spectra, both peptides 2 and 3 appear to exist as a mixture of conformers with the β‐hairpin as a minor conformer in aqueous solution, and a slight increase in hairpin population in 50% trifluoroethanol, which was more pronounced for peptide 3 . NMR spectra of peptide 2 in a 1:1 CD3CN/H2O mixture and 30 mM deuterated sodium dodecyl sulfate showed evidence of an extended backbone conformation of the β‐strand residues. However, inter‐strand rotating frame Overhauser effects (ROE) could not be detected and a loosely defined divergent hairpin structure resulted from ROE structure calculation in CD3CN/H2O. The loosely defined hairpin conformation is most likely a result of the electrostatic repulsions between cationic strand residues which also probably contribute towards maintaining low haemolytic activity.  相似文献   

14.
15.
16.
Carotenoid biosynthesis is highly conserved and well characterized up to the synthesis of β‐carotene. Conversely, the synthesis of astaxanthin from β‐carotene is less well characterized. Regardless, astaxanthin is a highly sought natural product, due to its various industrial applications and elevated antioxidant capacity. In this article, 12 β‐carotene ketolase and 4 β‐carotene hydroxylase genes, isolated from 5 cyanobacterial species, are investigated for their function, and potential for microbial astaxanthin synthesis. Further, this in vivo comparison identifies and applies the most promising genetic elements within a dual expression vector, which is maintained in Escherichia coli. Here, combined overexpression of individual β‐carotene ketolase and β‐carotene hydroxylase genes, within a β‐carotene accumulating host, enables a 23.5‐fold improvement in total carotenoid yield (1.99 mg g?1), over the parental strain, with >90% astaxanthin. Biotechnol. Bioeng. 2009;103: 944–955. © 2009 Wiley Periodicals, Inc.  相似文献   

17.
18.
Novel grafted agar disks were prepared for the covalent immobilization of β‐D‐galactosidase (β‐gal). The agar disks were activated through reacting with ethylenediamine or different molecular weights of Polyethyleneimine (PEI), followed by glutaraldehyde (GA). The modification of the agar gel and the binding of the enzyme were verified by Fourier Transform Infrared (FTIR) and elemental analysis. Moreover, the agar's activation process was optimized, and the amount of immobilized enzyme increased 3.44 folds, from 38.1 to 131.2 U/g gel, during the course of the optimization process. The immobilization of β‐gal onto the activated agar disks caused its optimum temperature to increase from 45°C to 45–55°C. The optimum pH of the enzyme was also shifted towards the acidic side (3.6–4.6) after its immobilization. Additionally, the Michaelis‐Menten constant (Km) increased for the immobilized β‐gal as compared to its free counterpart whereas the maximum reaction rate (Vmax) decreased. The immobilized enzyme was also shown to retain 92.99% of its initial activity after being used for 15 consecutive times. © 2015 Wiley Periodicals, Inc. Biopolymers 103: 675–684, 2015.  相似文献   

19.
A pseudo steady‐state model for the kinetically controlled synthesis of galacto‐oligosaccharides (GOS) with Aspergillus oryzae β‐galactosidase is presented. The model accounts for the dynamics of lactose consumption and production of galactose, glucose, di, tri, tetra, and penta‐oligosaccharides during the synthesis, being able to describe the total GOS content in the reaction medium at the experimental conditions evaluated. Experimental results show that the formation of GOS containing only galactose residues is significant at high conversions of substrate, which was taken into account in the model. The formation of enzyme transition complexes was considered and reasonable assumptions were made to reduce the number of parameters to be determined. The model developed has 8 parameters; 2 of them were experimentally determined and the other 6 were estimated by fitting to the experimental data using multiresponse regression. Temperature effect on kinetic and affinity constants was determined in the range from 40 to 55°C, and the data were fitted to Arrhenius type equation. Parameters of the proposed model are independent from the enzyme load in the reaction medium and, differently from previously reported models, they have a clear biochemical meaning. The magnitude of the kinetic and affinity constants of the enzyme suggests that the liberation of galactose from the galactosyl–enzyme complex is a very slow reaction and such complex is driven into GOS formation. It also suggests that the affinity for sugars of the galactosyl–enzyme complex is higher than that of the free enzyme. Biotechnol. Bioeng. 2011;108: 2270–2279. © 2011 Wiley Periodicals, Inc.  相似文献   

20.
Accurate and rapid detection of carbapenemases and identification of their types in Enterobacteriaceae are both still major challenges for clinical laboratories in attempting to prevent the intrusion and transmission of carbapenemase‐producing Enterobacteriaceae. This study aimed to evaluate the performance of the MASTDISCS combi Carba plus disc system in identification of different carbapenemase types, including OXA‐48‐type carbapenemase, for which no specific enzyme inhibitors have so far been available. The simple disc system discriminates carbapenemases, including OXA‐48‐types exhibiting low carbapenem minimum inhibitory concentrations, by targeting Enterobacteriaceae isolates with a EUCAST meropenem screening cut‐off of ≥0.25 mg/L.  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