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.
A fluorescence method was established for a α‐glucosidase activity assay and inhibitor screening based on β‐cyclodextrin‐coated quantum dots. p‐Nitrophenol, the hydrolysis product of the α‐glucosidase reaction, could quench the fluorescence of β‐cyclodextrin‐coated quantum dots via an electron transfer process, leading to fluorescence turn‐off, whereas the fluorescence of the system turned on in the presence of α‐glucosidase inhibitors. Taking advantage of the excellent properties of quantum dots, this method provided a very simple, rapid and sensitive screening method for α‐glucosidase inhibitors. Two α‐glucosidase inhibitors, 2,4,6‐tribromophenol and acarbose, were used to evaluate the feasibility of this screening model, and IC50 values of 24 μM and 0.55 mM were obtained respectively, which were lower than those previously reported. The method may have potential application in screening α‐glucosidase inhibitors. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

2.
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   

3.
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   

4.
Secondary metabolites and synthetic iminosugars that structurally resemble monosaccharides are potent inhibitors of α‐glucosidase activity. The enzyme is core in cleaving sucrose in phloem feeding insects and it also plays a crucial role of reducing osmotic stress via the formation of oligosaccharides. Inhibition of hydrolysis by iminosugars should result in nutritional deficiencies and/or disruption of normal osmoregulation. Deoxynojirimycin (DNJ) and 2 N‐alkylated analogs [N‐butyl DNJ (NB‐DNJ) and N‐nonyl DNJ (NN‐DNJ)] were the major iminosugars used throughout the study. The extensive experiments conducted with α‐glucosidase of the whitefly Bemisia tabaci indicated the competitive nature of inhibition and that the hydrophilic DNJ is a potent inhibitor in comparison to the more hydrophobic NB‐DNJ and NN‐DNJ compounds. The same inhibitory pattern was observed with the psyllid Cacopsylla bidens α‐glucosidase. In contrast to the above pattern, enzymes of the aphids, Myzus persicae and Aphis gossypii were more sensitive to the hydrophobic iminosugars as compared to DNJ. In vivo experiments in which adult B. tabaci were fed dietary iminosugars, show that the hydrophilic DNJ was far less toxic than the lipophilic NB‐DNJ and NN‐DNJ. It is proposed that this pattern is attributed to the better accessibility of the hydrophobic NN‐DNJ to the α‐glucosidase membrane‐bound compartment in the midgut. Based on the inhibitory effects of certain polyhydroxy N‐alkylated iminosugars, α‐glucosidase of phloem feeding hemipterans could serve as an attractive target site for developing novel pest control agents.  相似文献   

5.
Aims: To study glycosidase activities of a Lactobacillus brevis strain and to isolate an intracellular β‐glucosidase from this strain. Methods and Results: Lactic acid bacteria (LAB) isolated from a commercially available starter culture preparation for malolactic fermentation were tested for β‐glycosidase activities. A strain of Lact. brevis showing high intracellular β‐d ‐glucosidase, β‐d ‐xylosidase and α‐l ‐arabinosidase activities was selected for purification and characterization of its β‐glucosidase. The pure glucosidase from Lact. brevis has also side activities of xylosidase, arabinosidase and cellobiosidase. It is a homotetramer of 330 kDa and has an isoelectric point at pH 3·5. The Km for p‐nitrophenyl‐β‐d ‐glucopyranoside and p‐nitrophenyl‐β‐d ‐xylopyranoside is 0·22 and 1·14 mmol l?1, respectively. The β‐glucosidase activity was strongly inhibited by gluconic acid δ‐lactone, partially by glucose and gluconate, but not by fructose. Ethanol and methanol were found to increase the activity up to twofold. The free enzyme was stable at pH 7·0 (t1/2 = 50 day) but not at pH 4·0 (t1/2 = 4 days). Conclusions: The β‐glucosidase from Lact. brevis is widely different to that characterized from Lactobacillus casei ( Coulon et al. 1998 ) and Lactobacillus plantarum ( Sestelo et al. 2004 ). The high tolerance to fructose and ethanol, the low inhibitory effect of glucose on the enzyme activity and the good long‐term stability could be of great interest for the release of aroma compounds during winemaking. Significance and Impact of the study: Although the release of aroma compounds by LAB has been demonstrated by several authors, little information exists on the responsible enzymes. This study contains the first characterization of an intracellular β‐glucosidase isolated from a wine‐related strain of Lact. brevis.  相似文献   

6.
The potential to inhibit α‐ and β‐glucosidases of a series of chiral piperazine‐2,5‐dione derivatives was investigated. Three of the seven compounds tested, viz., 1, 5b , and 5c , showed to be non competitive inhibitors of α‐glucosidase, whereas they exhibited very low inhibitory activity towards β‐glucosidase. The most active compound, 5c (KI of α‐glucosidase=5 μm), had a 100‐fold α‐glucosidase/β‐glucosidase inhibitor selectivity.  相似文献   

7.
A link between senescence‐induced decline in photosynthesis and activity of β‐glucosidase is examined in the leaves of Arabidopsis. The enzyme is purified and characterized. The molecular weight of the enzyme is 58 kDa. It shows maximum activity at pH 5.5 and at temperature of 50°C. Photosynthetic measurements and activity of the enzyme are conducted at different developmental stages including senescence of leaves. Senescence causes a significant loss in total chlorophyll, stomatal conductance, rate of evaporation and in the ability of the leaves for carbon dioxide fixation. The process also brings about a decline in oxygen evolution, quantum yield of photosystem II (PS II) and quantum efficiency of PS II photochemistry of thylakoid membrane. The loss in photosynthesis is accompanied by a significant increase in the activity of the cell wall‐bound β‐glucosidase that breaks down polysaccharides to soluble sugars. The loss in photosynthesis as a signal for the enhancement in the activity of the enzyme is confirmed from the observation that incubation of excised mature leaves in continuous dark or in light with a photosynthesis inhibitor 3‐(3,4‐dichlorophenyl)‐1, 1‐dimethylurea (DCMU) that leads to sugar starvation enhances the activity of the enzyme. The work suggests that in the background of photosynthetic decline, the polysaccharides bound to cell wall that remains intact even during late phase of senescence may be the last target of senescing leaves for a possible source of sugar for remobilization and completion of the energy‐dependent senescence program.  相似文献   

8.
NagZ is an exo‐N‐acetyl‐β‐glucosaminidase, found within Gram‐negative bacteria, that acts in the peptidoglycan recycling pathway to cleave N‐acetylglucosamine residues off peptidoglycan fragments. This activity is required for resistance to cephalosporins mediated by inducible AmpC β‐lactamase. NagZ uses a catalytic mechanism involving a covalent glycosyl enzyme intermediate, unlike that of the human exo‐N‐acetyl‐β‐glucosaminidases: O‐GlcNAcase and the β‐hexosaminidase isoenzymes. These latter enzymes, which remove GlcNAc from glycoconjugates, use a neighboring‐group catalytic mechanism that proceeds through an oxazoline intermediate. Exploiting these mechanistic differences we previously developed 2‐N‐acyl derivatives of O‐(2‐acetamido‐2‐deoxy‐D ‐glucopyranosylidene)amino‐N‐phenylcarbamate (PUGNAc), which selectively inhibits NagZ over the functionally related human enzymes and attenuate antibiotic resistance in Gram‐negatives that harbor inducible AmpC. To understand the structural basis for the selectivity of these inhibitors for NagZ, we have determined its crystallographic structure in complex with N‐valeryl‐PUGNAc, the most selective known inhibitor of NagZ over both the human β‐hexosaminidases and O‐GlcNAcase. The selectivity stems from the five‐carbon acyl chain of N‐valeryl‐PUGNAc, which we found ordered within the enzyme active site. In contrast, a structure determination of a human O‐GlcNAcase homologue bound to a related inhibitor N‐butyryl‐PUGNAc, which bears a four‐carbon chain and is selective for both NagZ and O‐GlcNAcase over the human β‐hexosamnidases, reveals that this inhibitor induces several conformational changes in the active site of this O‐GlcNAcase homologue. A comparison of these complexes, and with the human β‐hexosaminidases, reveals how selectivity for NagZ can be engineered by altering the 2‐N‐acyl substituent of PUGNAc to develop inhibitors that repress AmpC mediated β‐lactam resistance.  相似文献   

9.
Metallo‐β‐lactamases (MBLs) are some of the best known β‐lactamases produced by common Gram‐positive and Gram‐negative pathogens and are crucial factors in the rise of bacterial resistance against β‐lactam antibiotics. Although many types of β‐lactamase inhibitors have been successfully developed and used in clinical settings, no MBL inhibitors have been identified to date. Nitrocefin, checkerboard and time‐kill assays were used to examine the enzyme behaviour in vitro. Molecular docking calculation, molecular dynamics simulation, calculation of the binding free energy and ligand‐residue interaction decomposition were used for mechanistic research. The behaviour of the enzymes in vivo was investigated by a mouse infection experiment. We showed that theaflavin‐3,3´‐digallate (TFDG), a natural compound lacking antibacterial activities, can inhibit the hydrolysis of MBLs. In the checkerboard and time‐kill assays, we observed a synergistic effect of TFDG with β‐lactam antibiotics against methicillin‐resistant Staphylococcus aureus BAA1717. Molecular dynamics simulations were used to identify the mechanism of the inhibition of MBLs by TFDG, and we observed that the hydrolysis activity of the MBLs was restricted by the binding of TFDG to Gln242 and Ser369. Furthermore, the combination of TFDG with β‐lactam antibiotics showed effective protection in a mouse Staphylococcus aureus pneumonia model. These findings suggest that TFDG can effectively inhibit the hydrolysis activity of MBLs and enhance the antibacterial activity of β‐lactam antibiotics against pathogens in vitro and in vivo.  相似文献   

10.
Carbohydrate hydrolyzing α‐glucosidases are commonly found in microorganisms present in the human intestine microbiome. We have previously reported crystal structures of an α‐glucosidase from the human gut bacterium Blaubia (Ruminococcus) obeum (Ro‐αG1) and its substrate preference/specificity switch. This novel member of the GH31 family is a structural homolog of human intestinal maltase‐glucoamylase (MGAM) and sucrase–isomaltase (SI) with a highly conserved active site that is predicted to be common in Ro‐αG1 homologs among other species that colonize the human gut. In this report, we present structures of Ro‐αG1 in complex with the antidiabetic α‐glucosidase inhibitors voglibose, miglitol, and acarbose and supporting binding data. The in vitro binding of these antidiabetic drugs to Ro‐αG1 suggests the potential for unintended in vivo crossreaction of the α‐glucosidase inhibitors to bacterial α‐glucosidases that are present in gut microorganism communities. Moreover, analysis of these drug‐bound enzyme structures could benefit further antidiabetic drug development.  相似文献   

11.
The commercial value of Terfezia claveryi, an edible desert truffle with important gastronomic, nutritional, and antioxidant properties, has led to growing interest in its cultivation. The erratic and slow growth of T. claveryi mycelium in vitro represents an impairment to obtain mycorrhizal plants, and it makes necessary to find a new culture medium able to overcome these drawbacks. In this work, we analyze the effect of cyclodextrins (CDs) on the growth of T. claveryi mycelium. Different parameters, including colony diameter, growth rate, and colony fresh weight, were evaluated, both in the presence and absence of these encapsulant agents. The results obtained confirm the ability of CDs to stimulate the growth of T. claveryi mycelium when present in the culture medium. A similar effect was observed when CDs were added to the culture medium of Tuber melanosporum. Three natural (α‐, β‐, and γ) and two modified (hydroxypropil‐β and methyl‐β) CDs were assayed. The best results were obtained with β‐cyclodextrin, but no improvement was observed with its chemically modified derivatives. CDs complex the different compounds present in the culture medium which impair mycelial growth. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 29:1558–1564, 2013  相似文献   

12.
Stem‐cell antigen 1–positive (Sca‐1+) cardiac stem cells (CSCs), a vital kind of CSCs in humans, promote cardiac repair in vivo and can differentiate to cardiomyocytes with 5′‐azacytizine treatment in vitro. However, the underlying molecular mechanisms are unknown. β‐arrestin2 is an important scaffold protein and highly expressed in the heart. To explore the function of β‐arrestin2 in Sca‐1+ CSC differentiation, we used β‐arrestin2–knockout mice and overexpression strategies. Real‐time PCR revealed that β‐arrestin2 promoted 5′‐azacytizine‐induced Sca‐1+ CSC differentiation in vitro. Because the microRNA 155 (miR‐155) may regulate β‐arrestin2 expression, we detected its role and relationship with β‐arrestin2 and glycogen synthase kinase 3 (GSK3β), another probable target of miR‐155. Real‐time PCR revealed that miR‐155, inhibited by β‐arrestin2, impaired 5′‐azacytizine‐induced Sca‐1+ CSC differentiation. On luciferase report assay, miR‐155 could inhibit the activity of β‐arrestin2 and GSK3β, which suggests a loop pathway between miR‐155 and β‐arrestin2. Furthermore, β‐arrestin2‐knockout inhibited the activity of GSK3β. Akt, the upstream inhibitor of GSK3β, was inhibited in β‐arrestin2‐Knockout mice, so the activity of GSK3β was regulated by β‐arrestin2 not Akt. We transplanted Sca‐1+ CSCs from β‐arrestin2‐knockout mice to mice with myocardial infarction and found similar protective functions as in wild‐type mice but impaired arterial elastance. Furthermore, low level of β‐arrestin2 agreed with decreased phosphorylation of AKT and increased phophorylation of GSK3β, similar to in vitro findings. The β‐arrestin2/miR‐155/GSK3β pathway may be a new mechanism with implications for treatment of heart disease.  相似文献   

13.
The greater wax moth, Galleria mellonella, is one of the most ruinous pests of honeycomb in the world. Beta‐glucosidases are a type of digestive enzymes that hydrolytically catalyzes the beta‐glycosidic linkage of glycosides. Characterization of the beta‐glucosidase in G. mellonella could be a significant stage for a better comprehending of its role and establishing a safe and effective control procedure primarily against G. mellonella and also some other insect pests. Laboratory reared final instar stage larvae were randomly selected and homogenized for beta‐glucosidase activity assay and subsequent analysis. The enzyme was purified to apparent homogeneity by salting out with ammonium sulfate and using sepharose‐4B‐l ‐tyrosine‐1‐naphthylamine hydrophobic interaction chromatography. The purification was 58‐fold with an overall enzyme yield of 29%. The molecular mass of the protein was estimated as ca. 42 kDa. The purified beta‐glucosidase was effectively active on para/ortho‐nitrophenyl‐beta‐d ‐glucopyranosides (p‐/o‐NPG) with Km values of 0.37 and 1.9 mM and Vmax values of 625 and 189 U/mg, respectively. It also exhibits different levels of activity against para‐nitrophenyl‐β‐d ‐fucopyranoside (p‐NPF), para/ortho‐nitrophenyl β‐d ‐galactopyranosides (p‐/o‐NPGal) and p‐nitrophenyl 1‐thio‐β‐d ‐glucopyranoside. The enzyme was competitively inhibited by beta‐gluconolactone and also was very tolerant to glucose against p‐NPG as substrate. The Ki and IC50 values of δ‐gluconolactone were determined as 0.021 and 0.08 mM while the enzyme was more tolerant to glucose inhibition with IC50 value of 213.13 mM for p‐NPG.  相似文献   

14.
Microbial β‐1,4‐galactanases are glycoside hydrolases belonging to family 53, which degrade galactan and arabinogalactan side chains in the hairy regions of pectin, a major plant cell wall component. They belong to the larger clan GH‐A of glycoside hydrolases, which cover many different poly‐ and oligosaccharidase specificities. Crystallographic complexes of Bacillus licheniformis β‐1,4‐galactanase and its inactive nucleophile mutant have been obtained with methyl‐β(1→4)‐galactotetraoside, providing, for the first time, information on substrate binding to the aglycone side of the β‐1,4‐galactanase substrate binding groove. Using the experimentally determined subsites as a starting point, a β(1→4)‐galactononaose was built into the structure and subjected to molecular dynamics simulations giving further insight into the residues involved in the binding of the polysaccharide from subsite ?4 to +5. In particular, this analysis newly identified a conserved β‐turn, which contributes to subsites ?2 to +3. This β‐turn is unique to family 53 β‐1,4‐galactanases among all clan GH‐A families that have been structurally characterized and thus might be a structural signature for endo‐β‐1,4‐galactanase specificity. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
β‐lactam antibiotics are crucial to the management of bacterial infections in the medical community. Due to overuse and misuse, clinically significant bacteria are now resistant to many commercially available antibiotics. The most widespread resistance mechanism to β‐lactams is the expression of β‐lactamase enzymes. To overcome β‐lactamase mediated resistance, inhibitors were designed to inactivate these enzymes. However, current inhibitors (clavulanic acid, tazobactam, and sulbactam) for β‐lactamases also contain the characteristic β‐lactam ring, making them susceptible to resistance mechanisms employed by bacteria. This presents a critical need for novel, non‐β‐lactam inhibitors that can circumvent these resistance mechanisms. The carbapenem‐hydrolyzing class D β‐lactamases (CHDLs) are of particular concern, given that they efficiently hydrolyze potent carbapenem antibiotics. Unfortunately, these enzymes are not inhibited by clinically available β‐lactamase inhibitors, nor are they effectively inhibited by the newest, non‐β‐lactam inhibitor, avibactam. Boronic acids are known transition state analog inhibitors of class A and C β‐lactamases, and are not extensively characterized as inhibitors of class D β‐lactamases. Importantly, boronic acids provide a novel way to potentially inhibit class D β‐lactamases. Sixteen boronic acids were selected and tested for inhibition of the CHDL OXA‐24/40. Several compounds were identified as effective inhibitors of OXA‐24/40, with Ki values as low as 5 μM. The X‐ray crystal structures of OXA‐24/40 in complex with BA3, BA4, BA8, and BA16 were determined and revealed the importance of interactions with hydrophobic residues Tyr112 and Trp115. These boronic acids serve as progenitors in optimization efforts of a novel series of inhibitors for class D β‐lactamases.  相似文献   

16.
NagZ is an N‐acetyl‐β‐d ‐glucosaminidase that participates in the peptidoglycan (PG) recycling pathway of Gram‐negative bacteria by removing N‐acetyl‐glucosamine (GlcNAc) from PG fragments that have been excised from the cell wall during growth. The 1,6‐anhydromuramoyl‐peptide products generated by NagZ activate β‐lactam resistance in many Gram‐negative bacteria by inducing the expression of AmpC β‐lactamase. Blocking NagZ activity can thereby suppress β‐lactam antibiotic resistance in these bacteria. The NagZ active site is dynamic and it accommodates distortion of the glycan substrate during catalysis using a mobile catalytic loop that carries a histidine residue which serves as the active site general acid/base catalyst. Here, we show that flexibility of this catalytic loop also accommodates structural differences in small molecule inhibitors of NagZ, which could be exploited to improve inhibitor specificity. X‐ray structures of NagZ bound to the potent yet non‐selective N‐acetyl‐β‐glucosaminidase inhibitor PUGNAc (O‐(2‐acetamido‐2‐deoxy‐d ‐glucopyranosylidene) amino‐N‐phenylcarbamate), and two NagZ‐selective inhibitors – EtBuPUG, a PUGNAc derivative bearing a 2‐N‐ethylbutyryl group, and MM‐156, a 3‐N‐butyryl trihydroxyazepane, revealed that the phenylcarbamate moiety of PUGNAc and EtBuPUG completely displaces the catalytic loop from the NagZ active site to yield a catalytically incompetent form of the enzyme. In contrast, the catalytic loop was found positioned in the catalytically active conformation within the NagZ active site when bound to MM‐156, which lacks the phenylcarbamate extension. Displacement of the catalytic loop by PUGNAc and its N‐acyl derivative EtBuPUG alters the active site conformation of NagZ, which presents an additional strategy to improve the potency and specificity of NagZ inhibitors.  相似文献   

17.
The aim of this study was to investigate the mechanism of the cytotoxic effect of β‐bungarotoxin (β‐BuTX), a presynaptic neurotoxin, on rat cerebellar granule neurons (CGNs). The maturation of CGNs is characterized by the prominent dense neurite networks that became fragmented after treatment with β‐BuTX, and this cytotoxic effect of β‐BuTX on CGNs was in a dose‐ and time‐dependant manner. The cytotoxic effect of β‐BuTX was found to be more potent than other toxins, such as α‐BuTX, cardiotoxin, melittin, and Naja naja atra venom phospholipase A2. Meanwhile, undifferentiated neuroblastoma neuronal cell lines, IMR‐32 and SK‐N‐MC, and astrocytes were found to be resistant to β‐BuTX. These results indicated that only the mature CGNs were sensitive to β‐BuTX insults. None of the following chemicals: antioxidants, K+‐channel activator, K+‐channel antagonists, intracellular Ca2+ chelator, Ca2+‐channel blockers, NMDA receptor antagonists, and nitric oxide synthase inhibitor tested, were able to reduce β‐BuTX‐induced cytotoxicity. However, secretory type phospholipase A2 inhibitors (glycyrrhizin and aristolochic acid) and a free radical scavenger (5,5‐dimethyl pyrroline N‐oxide, DMPO) could attenuate not only β‐BuTX‐induced cytotoxicity but also ROS production and caspase‐3 activation. These data suggest that phospholipase A2 activity of β‐BuTX may be responsible for free radical generation and caspase‐3 activation that accounts for the observed cytotoxic effect. It is proposed that the CGNs can be a useful tool for studying interactions of the molecules on neuronal plasma membrane with β‐BuTX that mediates the specific cytotoxicity. © 2005 Wiley Periodicals, Inc. J Neurobiol, 2005  相似文献   

18.
This article reports on a novel fluorescence resonance energy transfer (FRET) system between harmine and silver nanoparticles (AgNPs), in which harmine acts as the donor and AgNPs act as the acceptor. As a result of FRET, harmine fluorescence is quenched efficiently with a corresponding Stern–Volmer constant of 3.61 × 1011 L/mol. It was found that upon addition of the anticancer drug, 6–thioguanine (6–TG), the fluorescence was recovered due to the competitive adsorption of this compound onto AgNPs. Based on this effect, a selective turn‐on fluorescence sensor was developed for the determination of 6–TG. Under optimum conditions, the enhanced fluorescence intensity displays a linear relationship with the concentration of 6–TG in the range 1.5 × 10‐8–7.5 × 10‐7 M with a detection limit of 9.7 nM. The developed method was applied to the determination of this drug in a pharmaceutical preparation and human plasma samples. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

19.
Phosphoinositide 3‐kinase (PI3K) signaling is frequently deregulated in breast cancer and plays a critical role in tumor progression. However, resistance to PI3K inhibitors in breast cancer has emerged, which is due to the enhanced β‐catenin nuclear accumulation. Until now, the mechanisms underlying PI3K inhibition‐induced β‐catenin nuclear accumulation remains largely unknown. In the present study, we found inhibition of PI3K with LY294002 promoted β‐catenin nuclear accumulation in MCF‐7 and MDA‐MB‐231 breast cancer cells. Combining PI3K inhibitor LY294002 with XAV‐939, an inhibitor against β‐catenin nuclear accumulation, produced an additive anti‐proliferation effect against breast cancer cells. Subsequent experiments suggested β‐catenin nuclear accumulation induced by PI3K inhibition depended on the feedback activation of epidermal growth factor receptor (EGFR) signaling pathway in breast cancer cells. Inhibition of EGFR phosphorylation with Gefitinib enhanced anti‐proliferation effect of PI3K inhibitor LY294002 in MCF‐7 and MDA‐MB‐231 cells. Taken together, our findings may elucidate a possible mechanism explaining the poor outcome of PI3K inhibitors in breast cancer treatment.  相似文献   

20.
We established a new animal model called SPORTS (Spontaneously‐Running Tokushima‐Shikoku) rats, which show high‐epinephrine (Epi) levels. Recent reports show that Epi activates adenosine monophosphate (AMP)–activated protein kinase (AMPK) in adipocytes. Acetyl‐CoA carboxylase (ACC) is the rate‐limiting enzyme in fatty acid synthesis, and the enzymatic activity is suppressed when its Ser‐79 is phosphorylated by AMPK. The aim of this study was to investigate the in vivo effect of Epi on ACC and abdominal visceral fat accumulation. We divided both 6‐week male control and SPORTS rats into two groups, which were fed either normal diet or high fat and sucrose (HFS) diet for 16 weeks. At the end of diet treatment, retroperitoneal fat was collected for western blotting and histological analysis. Food intake was not different among the groups, but SPORTS rats showed significantly lower weight gain than control rats in both diet groups. After 10 weeks of diet treatment, glucose tolerance tests (GTTs) revealed that SPORTS rats had increased insulin sensitivity. Furthermore, SPORTS rats had lower quantities of both abdominal fat and plasma triglyceride (TG). In abdominal fat, elevated ACC Ser‐79 phosphorylation was observed in SPORTS rats and suppressed by an antagonist of β‐adrenergic receptor (AR), propranolol, or an inhibitor of AMPK, Compound C. From these results, high level of Epi induced ACC phosphorylation mediated through β‐AR and AMPK signaling pathways in abdominal visceral fat of SPORTS rats, which may contribute to reduce abdominal visceral fat accumulation and increase insulin sensitivity. Our results suggest that β‐AR‐regulated ACC activity would be a target for treating lifestyle‐related diseases, such as obesity.  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