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.
Phospholipase A(2) (PLA(2)) quantitation in real-time, using (7-nitro-2-1,3-benzoxadiazol-4-yl)amino-derivatives of phosphatidylcholine (NBD-PCs) as substrates, is influenced by high protein content, color or turbidity. The purpose of the study was to overcome such limitations by a complementary reversed-phase HPLC step to separate the substrates from the products of the reaction. Plasma and bronchoalveolar lavage (BAL) fluid were applied as enzymic sources, while standard porcine PLA(2) and human plasma PAF-acetylhydrolase (PAF-AH) were employed as positive controls. The method was validated with a radiometric assay and compared with the real-time fluorimetric assay. Regarding PLA(2) and PAF-AH determination, the isocratic elution systems CH(3)OH-H(2)O (80:20, v/v) and CH(3)OH-H(2)O-CH(3)COOH (60:40:0.2, v/v/v) separated efficiently C(12)-NBD-FA/C(12)-NBD-PC and C(6)-NBD-FA/C(6)-NBD-PC, with 4.4 and 2.2 resolution, respectively. Analysis time was ~15 min/injection. The reproducibility, expressed as relative standard deviation, was ≤13% for PLA(2) and ≤16% for PAF-AH, respectively. The assay was linear for PLA(2) activities extending from 1 pmol up to at least 250 nmol FA/h/mL sample, similar to the radiometric assay. It was appropriate for samples with high protein content, where the real-time fluorimetric assay was insufficient. The HPLC method was also evaluated under elevated temperatures, various pH values and Ca(2+) concentrations.  相似文献   

2.
The interfacial activation of porcine pancreatic phospholipase A(2) (PLA(2)) during the hydrolysis of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine liposomes at different temperatures has been monitored by fluorescence changes of the 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) lipid derivatives 1-palmitoyl-2-[6-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-glycero-3-phosphocholine (C(12)-NBD-PC) and 12-[(7-nitrobenz-2-oxa-1,3-diazol-4-yl)amino)]dodecanoic acid (C(12)-NBD-FA) inserted in the substrate vesicles. These long-chain monitors, in contrast to the previously used C(6)-NBD-PC, detect latency times of PLA(2) action, similar to those measured by the classic titrimetric, pH-stat method. Interestingly, hydrolysis of the host vesicles results in a decrease in fluorescence not only of C(12)-NBD-PC, a substrate analog, but also of product derivative C(12)-NBD-FA. Ultrafiltration experiments show that C(12)-NBD-FA does not migrate to the aqueous phase upon hydrolysis of the host liposomes. Besides, in a simulated hydrolysis experiment in which increasing proportions of palmitic acid and 1-palmitoyl-sn-glycero-3-phosphocholine were cosonicated with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, C(12)-NBD-PC fluorescence was insensitive to products, whereas C(12)-NBD-FA did show a decreased emission intensity as in the actual hydrolysis experiments. The phenomenon is triggered above a critical concentration of products (10 mol%) suggesting that cosegregation of NBD-FA (either added as such or generated by hydrolysis of C(12)-NBD-PC) and products may be related to the decrease in fluorescence. Phase separation should create microdomains of increased C(12)-NBD-FA surface density and cause concentration quenching. In addition, and taking into account that the NBD group may be located near the interfacial region, it is possible that in segregating with products, the fluorescent moiety of C(12)-NBD-FA becomes exposed to microenvironments of higher surface polarity, which further decreases its quantum yield.  相似文献   

3.
We have examined the internalization and degradation of a fluorescent analog of phosphatidylcholine after its insertion into the plasma membrane of cultured Chinese hamster fibroblasts. 1-acyl-2-(N-4- nitrobenzo-2-oxa-1,3-diazole)-aminocaproyl phosphatidylcholine (C6-NBD- PC) was incorporated into the cell surface by liposome-cell lipid transfer at 2 degrees C. The fluorescent lipid remained localized at the plasma membrane as long as the cells were kept at 2 degrees C; however, when the cells were warmed to 37 degrees C, internalization of some of the fluorescent lipid occurred. Most of the internalized C6-NBD- PC accumulated in the Golgi apparatus although a small amount was found randomly distributed throughout the cytoplasm in punctate fluorescent structures. Internalization of the fluorescent lipid at 37 degrees C was blocked by the presence of inhibitors of endocytosis. Incubation of cells containing C6-NBD-PC at 37 degrees C resulted in a rapid degradation of the fluorescent lipid. This degradation occurred predominantly at the plasma membrane. The degradation of C6-NBD-PC resulted in the release of NBD-fatty acid into the medium. We have compared the internalization of the fluorescent lipid with that of a fluorescent protein bound to the cell surface. Both fluorescent lipid and protein remained at the plasma membrane at 2 degrees C and neither were internalized at 37 degrees C in the presence of inhibitors of endocytosis. However, when incubated at 37 degrees C under conditions that permit endocytosis, the two fluorescent species appeared at different intracellular sites. Our data suggest that there is no transmembrane movement of C6-NBD-PC and that the fluorescent probe reflects the internalization of the outer leaflet of the plasma membrane lipid bilayer. The results are consistent with the Golgi apparatus as being the primary delivery site of phospholipid by bulk membrane movement from the plasma membrane.  相似文献   

4.
We present a combined theoretical (molecular dynamics, MD) and experimental (differential scanning calorimetry, DSC) study of the effect of 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) acyl chain-labeled fluorescent phospholipid analogs (C6-NBD-PC and C12-NBD-PC) on 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) bilayers. DSC measurements reveal that < 1 mol% of NBD-PC causes elimination of the pre-transition and a large loss of cooperativity of the main transition of DPPC. Labeling with C6-NBD-PC or C12-NBD-PC shifts the main transition temperature to lower or higher values, respectively. Following our recent report on the location and dynamics of these probes (BBA 1768 (2007) 467-478) in fluid phase DPPC, we present a detailed analysis of 100-ns MD simulations of systems containing either C6-NBD-PC or C12-NBD-PC, focused on their influence on several properties of the host bilayer. Whereas most monitored parameters are not severely affected for 1.6 mol% of probe, for the higher concentration studied (6.2 mol%) important differences are evident. In agreement with published reports, we observed that the average area per phospholipid molecule increases, whereas DPPC acyl chain order parameters decrease. Moreover, we predict that incorporation of NBD-PC should increase the electrostatic potential across the bilayer and, especially for C12-NBD-PC, slow lateral diffusion of DPPC molecules and rotational mobility of DPPC acyl chains.  相似文献   

5.
A continuous fluorescence assay was adapted to the measurement of the phospholipase reaction of lecithin cholesterol acyltransferase (LCAT). The fluorescent phospholipid 1-acyl-2-(N-4-nitrobenzo-2-oxa-1,3-diazole)aminocaproyl phosphatidylcholine (C6-NBD-PC) in micelle form reacted with LCAT to yield NBD-caproic acid, resulting in up to 5-fold increases in fluorescence in 30 min. The reaction rates were optimal in mixtures containing 0.1 M NaCl and 4 mM beta-mercaptoethanol at 37 degrees C. Apolipoprotein A-I did not activate the enzyme and bovine serum albumin bound monomeric substrate and interfered with the fluorescence assay. Under similar reaction conditions, bee venom phospholipase A2 was almost 100-fold more reactive than LCAT.  相似文献   

6.
Location and dynamic reorientation of the fluorophore 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) covalently attached to a short (C6) or a long (C12) sn2 acyl chain of a phosphatidylcholine molecule was investigated by fluorescence and solid-state NMR spectroscopy. 2H NMR lipid chain order parameters indicate a perturbation of the phospholipid packing density in the presence of NBD. Specifically, a decrease of molecular order was found for acyl chain segments of the lower, more hydrophobic region. Molecular collision probabilities determined by 1H magic angle spinning nuclear Overhauser enhancement spectroscopy indicate a highly dynamic reorientation of the probe in the membrane due to thermal fluctuations. A broad distribution of the fluorophore in the lipid bilayer is observed with a preferential location in the upper acyl chain/glycerol region. The distribution of the NBD group in the membrane is quite similar for both the long- and the short-chain analog. However, a slight preference of the NBD group for the lipid-water interface is found for C12-NBD-PC in comparison with C6-NBD-PC. Indeed, as shown by dithionite fluorescence assay, the long-chain analog reacts more favorably with dithionite, indicating a better accessibility of the probe by dithionite present in the aqueous phase. Forces determining the location of the fluorophore in the lipid water interface are discussed.  相似文献   

7.
Platelet-activating factor (PAF)-acetylhydrolase is the enzyme modulating in tissues and biological fluids the concentration of the proinflammatory factors PAF and PAF-like oxidation products of phospholipids (PAF-like compounds). We investigated whether there is a relation between PAF-acetylhydrolase activity and the concentration of PAF-like compounds in bronchoalveolar lavage (BAL). We found that alveolar type II cells are an additional source of PAF-acetylhydrolase in BAL beside macrophages. Secretion of PAF-acetylhydrolase was stimulated by phorbol ester in alveolar type II cells but not in macrophages. Studies in BAL suggested that secreted PAF-acetylhydrolase was bound to alveolar surfactant. Exposure of rats to high oxygen concentration reduced the activity of PAF-acetylhydrolase in BAL and macrophages, but not in plasma or alveolar type II cells. In contrast, hyperoxia increased the concentration of PAF-like-compounds, lipid hydroperoxides and malonedialdehyde in plasma but not in BAL. Therefore, we conclude that neither the oxidant-induced decrease of the PAF-acetylhydrolase activity nor the direct peroxidation of surfactant lipids in the alveoli provide a likely mechanism for hyperoxia-induced lung injury. Instead, lung injury is apparently caused by lipid peroxidation in plasma rather than by high oxygen pressure in the alveoli.  相似文献   

8.
The fluorescent phospholipid analog N-(lissamine rhodamine B sulfonyl)phosphatidylethanolamine (N-Rh-PE) was inserted into the plasma membrane of Baby hamster kidney cells at low temperature (2 degrees C). The mobility characteristics of the analog--as revealed by fluorescence photobleaching recovery--were very similar to those of membrane-inserted 1-acyl-2[6-[N-(7-nitro-2,1,3-benzoxadiazol-4-yl) amino]caproyl] phosphatidylcholine (C6-NBD-PC). Upon warming to 37 degrees C, followed by a 1-h incubation, all N-Rh-PE was located intracellularly. By contrast, after the same time interval, approximately 10% of the cell-associated PC-derivative was found intracellularly. Furthermore, the analogs moved to different intracellular sites, as N-Rh-PE associates with perinuclear and peri-Golgi structures, whereas C6-NBD-PC appears mainly in the Golgi complex. Colocalization with organelle-specific probes and Percoll gradient analysis identified the N-Rh-PE-labeled structures as lysosomes. Temperature and energy-dependent experiments supported the endocytic pathway as the mechanism of N-Rh-PE internalization. The mechanism of N-Rh-PE internalization appears to differ from that of C6-NBD-PC. In conjunction with a difference in the efficiency of removal of the lipid derivatives from the plasma membrane, the results suggest that N-Rh-PE is selectively internalized, implying that sorting of the lipid analogs already occurs at the level of the plasma membrane. The distinct difference in physical appearance of the probes after membrane insertion, i.e., N-Rh-PE being present as small clusters and C6-NBD-PC as monomers, could explain the selective sorting and internalization of N-Rh-PE. The results demonstrate that N-Rh-PE may serve as a useful marker for studying membrane traffic during endocytosis.  相似文献   

9.
We have developed two high-performance liquid chromatography methods for separating a number of fluorescent 4-nitrobenzo-2-oxa-1,3-diazole (NBD) analogs of glycerolipids and sphingolipids. Samples of fluorescent lipid analogs containing NBD-aminocaproyl (C6-NBD) or NBD-aminododecanoyl (C12-NBD) acyl chains were synthesized and analyzed by the following HPLC methods. An isocratic normal-phase method permitted resolution of a mixture of the 1,2-(palmitoyl, C6-NBD)-analogs of triacylglycerol, diacylglycerol, phosphatidic acid, phosphatidylethanolamine, and phosphatidylcholine in less than 10 min, while a mixture of the (C6-NBD)-labeled analogs of ceramide, glucocerebroside, and sphingomyelin was separated in approximately 15 min. This method also detected various (C6-NBD)-phosphatidylcholine and -phosphatidylethanolamine molecules which differed only in their nonfluorescent acyl (oleoyl or palmitoyl) chains, and readily separated nonfluorescent dipalmitoylphosphatidylcholine from both (C6-NBD)- and (C12-NBD)-phosphatidylcholine derivatives. An isocratic reverse-phase system permitted separation of isomers of fluorescent phosphatidylcholine, -ethanolamine, -glycerol, -inositol, -serine, and phosphatidic acid in which the NBD-fatty acid was present in either the sn-1 or sn-2 position of the glycerol backbone.  相似文献   

10.
To investigate the effect of fluorescent probe on the properties of membranes, we studied model membranes composed of 1,2- dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and 1-palmitoyl 2-oleoyl-sn-glycero-3-phosphocholine (POPC) in the presence and absence of fluorescent probe. The morphology of giant unilamellar vesicles (GUVs) has been observed as a function of temperature and composition by fluorescence microscopy using NBD-DOPE or C6-NBD-PC as the probe. The phase behavior of model membranes containing no fluorescent probe was investigated by 2H-NMR spectroscopy. We found that the bright phase observed on GUVs was the fluid phase enriched in POPC and the dark phase was the gel phase enriched in DPPC. NBD-DOPE and C6-NBD-PC preferentially participated in the fluid-phase domains when GUVs were in the gel?+?fluid phase coexistence. Inclusion of both fluorescent probes (1?mol%) lowered the transition temperature of POPC/DPPC membranes. In addition, C6-NBD-PC exhibited a stronger effect than NBD-DOPE, which was considered to be associated with the structures of fluorescent molecules.  相似文献   

11.
100-ns molecular dynamics simulations of fluid 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) bilayers, both pure and containing 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) acyl-chain labeled fluorescent analogs (C6-NBD-PC and C12-NBD-PC), are described. These molecules are widely used as probes for lipid structure and dynamics. The results obtained here for pure DPPC agree with both experimental and theoretical published works. We verified that the NBD fluorophore of both derivatives loops to a transverse location closer to the interface than to the center of the bilayer. Whereas this was observed previously in experimental literature works, conflicting transverse locations were proposed for the NBD group. According to our results, the maximum of the transverse distribution of NBD is located around the glycerol backbone/carbonyl region, and the nitro group is the most external part of the fluorophore. Hydrogen bonds from the NH group of NBD (mostly to glycerol backbone lipid O atoms) and to the nitro O atoms of NBD (from water OH groups) are continuously observed. Rotation of NBD occurs with approximately 2.5-5 ns average correlation time for these probes, but very fast, unresolved reorientation motions occur in <20 ps, in agreement with time-resolved fluorescence anisotropy measurements. Finally, within the uncertainty of the analysis, both probes show lateral diffusion dynamics identical to DPPC.  相似文献   

12.
The kinetics of poly(ethylene glycol) (PEG)-induced fusion between intact human erythrocytes was continuously monitored by a fluorescence lipid mixing method, utilizing the dequenching of the fluorescence probe, 1-oleoyl-2-[12-[(7-nitro-2,1,3-benzoxadiazol-4-yl)amino]dodecanoyl ] phosphatidylcholine (C12-NBD-PC). The steady-state fluorescence intensity was detected from the surface of cells in a monolayer on an alcian blue-coated glass coverslip. The relief of fluorescence self-quenching after fusion between C12-NBD-PC labeled and unlabeled intact erythrocytes was measured. The extent of fluorescence dequenching was normalized based on the measured concentration of probes in membranes, the projected partial dequenching due both to dilution by intercellular fusion, and the dilution between the inner and outer leaflets of membranes (flip-flop). There was no significant increase in fluorescence intensity during PEG treatment of 5 min, at 4 degrees C. Intensity increased immediately after the dilution of PEG, and reached saturation in 30 min. The efficiency of fusion increased with the increasing of PEG concentrations. Only 4% enhancement of saturated relative fluorescence intensity was detected in 25 wt% PEG-induced cell fusion; 23% enhancement in 30 wt%; and 66% enhancement in 35 wt%. The transfer of fluorescent probes between membrane bilayer leaflets (flip-flop) was also monitored during the fusion process. Flip-flop was monitored in confluent monolayers as well as in isolated cells. There was no significant spontaneous flip-flop within 30 min of dilution. The relative fluorescence intensity enhancement contributed by the dilution of probes between fused labeled and unlabeled cells (at a 1:1 ratio) was found to account for only 39% of the observed final dequenching, whereas the contribution by flip-flop associated with cell fusion was found to account for 9%, and flip-flop without fusion contributed approximately 18%. A portion of the flip-flop is a consequence of hemolysis. Therefore, fluorescence dequenching measurements of fusion of whole cells must be interpreted with caution.  相似文献   

13.
The spectroscopic and ionization properties of various lipids labeled with the 7-nitrobenz-2-oxa-1,3-diazol-4-yl (NBD) group have been studied in model membranes using fluorescence, absorbance and electrophoretic mobility measurements. Electrophoretic measurements show that the NBD group is uncharged at neutral pH. However, at high pH, hydroxyl addition or deprotonation occurs with a pKa, depending upon conditions, of 11.5-11.8 for the NBD group of headgroup-labeled phosphatidylethanolamine (NBD-PE) and 11.1-11.5 for NBD labels placed at the end of one fatty acyl chain of a phosphatidylcholine (6-NBD-PC and 12-NBD-PC). This type of behavior is not observed in the case of a methylated NBD label placed in the flexible 'tail' of cholesterol (NBD-cholesterol). The similarity in pKa for NBD-PE and NBD-PCs suggests that in these cases the NBD group is at a similar depth in the membrane. This was examined further by comparison of the fluorescence emission maximum of the NBD group in model membranes with that in solvents of varying polarity. The apparent polarity experienced by NBD groups in model membranes indicates that for NBD-PE and 12-NBD-PC they are located at the polar region whereas the NBD group of NBD-cholesterol is deeply buried in a nonpolar region of the membrane. This conclusion is supported further by fluorescence quenching experiments measuring NBD exposure to the aqueous quencher Co2+. The results of this study confirm the tentative conclusions of our previous fluorescence quenching studies on the location of NBD groups in model membranes.  相似文献   

14.
The internalization and distribution of fluorescent analogs of phosphatidylcholine (M-C6-NBD-PC) and phosphatidylethanolamine (M-C6-NBD-PE) were studied in Saccharomyces cerevisiae. At normal growth temperatures, M-C6-NBD-PC was internalized predominantly to the vacuole and degraded. M-C6-NBD-PE was internalized to the nuclear envelope/ER and mitochondria, was not transported to the vacuole, and was not degraded. At 2 degrees C, both were internalized to the nuclear envelope/ER and mitochondria by an energy-dependent, N-ethylmaleimide-sensitive process, and transport of M-C6-NBD-PC to and degradation in the vacuole was blocked. Internalization of neither phospholipid was reduced in the endocytosis-defective mutant, end4-1. However, following pre-incubation at 37 degrees C, internalization of both phospholipids was inhibited at 2 degrees C and 37 degrees C in sec mutants defective in vesicular traffic. The sec18/NSF mutation was unique among the sec mutations in further blocking M-C6-NBD-PC translocation to the vacuole suggesting a dependence on membrane fusion. Based on these and previous observations, we propose that M-C6-NBD-PC and M-C6-NBD-PE are transported across the plasma membrane to the cytosolic leaflet by a protein-mediated, energy-dependent mechanism. From the cytosolic leaflet, both phospholipids are spontaneously distributed to the nuclear envelope/ER and mitochondria. Subsequently, M-C6-NBD-PC, but not M-C6-NBD-PE, is sorted by vesicular transport to the vacuole where it is degraded by lumenal hydrolases.  相似文献   

15.
The tescalcin gene is preferentially expressed during mouse testis differentiation. Here, we demonstrate that this gene encodes a 24 kDa Ca(2+)- and Mg(2+)-binding protein with one consensus EF-hand and three additional domains with EF-hand homology. Equilibrium dialysis with (45)Ca(2+) revealed that recombinant tescalcin binds approximately one Ca(2+) ion at physiological concentrations (pCa 4.5). The intrinsic tryptophan fluorescence of tescalcin was significantly reduced by Ca(2+), indicative of a conformational change. The apparent K(d) for Ca(2+) was 0.8 microM. A point mutation in the consensus EF-hand (D123A) abolished (45)Ca(2+) binding and prevented the fluorescence quenching, demonstrating that the consensus EF-hand alone mediates the Ca(2+)-induced conformational change. Tescalcin also binds Mg(2+) (K(d) 73 microM), resulting in a much smaller fluorescence decrease. In the presence of 1 mM Mg(2+), tescalcin's Ca(2+) affinity is shifted to 3.5 microM. These results illustrate that tescalcin should bind Mg(2+) constitutively in a quiescent cell, replacing it with Ca(2+) during stimulation. We also show that tescalcin is most abundant in adult mouse heart, brain, and stomach, as well as in HeLa and HL-60 cells. Immunofluorescence microscopy revealed that tescalcin is present in the cytoplasm and nucleus, with concentration in membrane ruffles and lamellipodia in the presence of serum, where it colocalizes with the small guanosine triphosphatase Rac-1. Tescalcin shares sequence and functional homology with calcineurin-B homologous protein (CHP), and we found that tescalcin, like CHP, can inhibit the phosphatase activity of calcineurin A. Hence, tescalcin is a novel calcineurin B-like protein that binds a single Ca(2+) ion.  相似文献   

16.
Oxidative stress may be a key feature, and hence important determinant, of tissue injury and allograft rejection in lung transplant recipients. To investigate this, we determined the antioxidant status (urate, ascorbate, thiols and alpha-tocopherol) and lipid peroxidation status (malondialdehyde) in bronchoalveolar lavage (BAL) fluid and blood serum of 19 consecutive lung transplant recipients 2 weeks and 1, 2, 3, 6, and 12 months post-surgery. BAL fluid and blood samples from 23 control subjects and blood from 8 patients two days before transplantation were obtained for comparison. Before surgery, the antioxidant status of patients was poor as serum ascorbate and total thiol concentrations were significantly (p < 0.05) lower than control subjects. Two weeks post-surgery, ascorbate and total thiol concentrations were still low and urate concentrations had fallen compared to control subjects (p < 0.01). At this time, BAL fluid urate concentration was higher (p < 0.01), ascorbate concentration was lower (p < 0.01) and reduced glutathione concentrations were similar to control subjects. MDA, a product of lipid peroxidation, was higher (p < 0.01) in both BAL fluid and serum obtained from transplant patients compared to control subjects. During the first 12 months post-surgery, little improvement in antioxidant status or extent of lipid peroxidation was seen in transplant recipients. Regression analysis indicated no difference in serum or BAL fluid antioxidant status in patients with acute rejection compared to those without. In conclusion, lung transplant recipients have a compromised antioxidant status before surgery and it remains poor for at least the first year following the operation. In addition, these patients have elevated MDA concentrations in both their lung lining fluid and blood over most of this time. Oxidative stress is not, however, a sufficiently sensitive endpoint to predict tissue rejection in this group.  相似文献   

17.
Studies have shown diabetes to be associated with alterations in composition of extracellular matrix and that such proteins modulate signal transduction. The present studies examined if non-enzymatic glycation of fibronectin or a mixed matrix preparation (EHS) alters endothelial cell Ca(2+) signaling following agonist stimulation. Endothelial cells were cultured from bovine aorta and rat heart. To glycate proteins, fibronectin (10 microg/ml), or EHS (2.5 mg/ml) were incubated (37 degrees C, 30 days) with 0.5 M glucose-6-phosphate. Matrix proteins were coated onto cover slips after which cells (10(5) cells/ml) were plated and allowed to adhere for 16 h. For measurement of intracellular Ca(2+), cells were loaded with fura 2 (2 microM) and fluorescence intensity monitored. Bovine cells on glycated EHS showed decreased ability for either ATP (10(-6) M) or bradykinin (10(-7) M) to increase Ca(2+) (i). In contrast, glycated fibronectin did not impair agonist-induced increases in Ca(2+) (i). In the absence of extracellular Ca(2+), ATP elicited a transient increase in Ca(2+) (i) consistent with intracellular release. Re-addition of Ca(2+) resulted in a secondary rise in Ca(2+) (i) indicative of store depletion-mediated Ca(2+) entry. Both phases of Ca(2+) mobilization were reduced in cells on glycated mixed matrix; however, as the ratio of the two components was similar in all cells, glycation appeared to selectively impair Ca(2+) release from intracellular stores. Thapsigargin treatment demonstrated an impaired ability of cells on glycated EHS to increase cytoplasmic Ca(2+) consistent with decreased endoplasmic reticulum Ca(2+) stores. Further support for Ca(2+) mobilization was provided by increased baseline IP(3) levels in cells plated on glycated EHS. Impaired ATP-induced Ca(2+) release could be induced by treating native EHS with laminin antibody or exposing cells to H(2)O(2) (20-200 microM). Glycated EHS impaired Ca(2+) signaling was attenuated by treatment with aminoguanidine or the antioxidant alpha-lipoic acid. The results demonstrate that matrix glycation impairs agonist-induced Ca(2+) (i) increases which may impact on regulatory functions of the endothelium and implicate possible involvement of oxidative stress.  相似文献   

18.
In this study emission and synchronous-scan fluorescence spectroscopy have been used to investigate the interaction of the class A (oxygen seeking 'hard acid') metal Al(3+), with Suwannee River fulvic acid (SRFA), as well as competition between Al(3+) and several other metal ions (Ca(2+), Mg(2+), Cu(2+), Pd(2+), La(3+), Tb(3+) and Fe(3+)) for binding sites on SRFA. Of the four metal ions possessing very similar (and relatively low) ionic indices (Ca(2+), Mg(2+), Cu(2+) and Pd(2+)) only the latter two paramagnetic ions significantly quenched SRFA fluorescence emission intensity. Of the four metal ions possessing very similar (and relatively low) covalent indices (Ca(2+), Mg(2+), La(3+) and Tb(3+)) only the last paramagnetic ion (Tb(3+)) significantly quenched SRFA fluorescence. None of these metals was able to significantly compete with SRFA-bound Al(3+).Fe(3+), which differs substantially from all of the other metals examined in this study in that it possesses a relatively high ionic index (but not as high as Al(3+)) and a relatively low covalent index (but not as low as Al(3+)), was able not only to quench SRFA fluorescence but also to compete (at least to some extent) with SRFA-bound Al(3+). Synchronous-scan fluorescence SRFA spectra taken in the absence and presence of Fe(3+) and/or Al(3+) support the view that these two metal ions can compete for sites on SRFA. The results of these fluorescence experiments further confirm the Al(3+), and metal ions that have electronic properties somewhat similar to Al(3+) (such as Fe(3+)) are somewhat unique in their ability to interact strongly with binding sites on fulvic acids.  相似文献   

19.
Marine bacterium Vibrio sp. F-6, utilizing agarose as a carbon source to produce agarases, was isolated from seawater samples taken from Qingdao, China. Two agarases (AG-a and AG-b) were purified to a homogeneity from the cultural supernatant of Vibrio sp. F-6 through ammonium sulfate precipitation, Q-Sepharose FF chromatography, and Sephacryl S-100 gel filtration. Molecular weights of agarases were estimated to be 54.0 kDa (AG-a) and 34.5 kDa (AG-b) by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). The optimum pH values for AG-a and AG-b were about 7.0 and 9.0, respectively. AG-a was stable in the pH range of 4.0-9.0 and AG-b was stable in the pH range of 4.0-10.0. The optimum temperatures of AG-a and AG-b were 40 and 55 degrees C, respectively. AG-a was stable at temperature below 50 degrees C. AG-b was stable at temperature below 60 degrees C. Zn(2+), Mg(2+) or Ca(2+) increased AG-a activity, while Mn(2+), Cu(2+) or Ca(2+) increased AG-b activity. However, Ag(+), Hg(2+), Fe(3+), EDTA and SDS inhibited AG-a and AG-b activities. The main hydrolysates of agarose by AG-a were neoagarotetraose and neoagarohexaose. The main hydrolysates of agarose by AG-b were neoagarooctaose and neoagarohexaose. When the mixture of AG-a and AG-b were used, agarose was mainly degraded into neoagarobiose.  相似文献   

20.
Five groups of ten female guinea-pigs were passively sensitized against ovalbumin (OA) (n = 9) or control guinea-pig serum (n = 1). 24 h later, they received mepyramine (0.5 mg/kg, i.p.) and 30 min later inhaled aerosols of: (A) OA (2 in 0.9% saline, 8 min, n = 4/9); (B) saline (40 min, n = 4/9); (C) LPS (40 min, Escherichia coli 0111:B4, 150 ng/kg in PBS, n = 1/9); and (D) the control animal was treated as in (C) (n = 1). Their tracheas were cannulated under pentobarbital anaesthesia and bronchial alveolar lavage (BAL) was performed with 2 x 5 ml PBS containing BSA (1%) (n = 1 group), or BSA (1%) and aprotinin (1000 KIU/ml) (n = 4 groups), at 30, 60, 90 or 120 min post-inhalations. BAL fluids recovered were centrifuged, the supernatants recovered and frozen until assayed for tumour necrosis factor-alpha (TNF-alpha), interleukin-1 (IL-1) and interleukin-6 (IL-6). No TNF-alpha could be detected unless aprotinin was present in the lavaging solution. BAL fluid from OA-sensitized and control animals that had inhaled LPS contained high levels of TNF-alpha that peaked at 90 min. BAL fluid from OA sensitized animals that inhaled OA aerosols contained no detectable TNF-alpha at 30 min, but it was found in increasing amounts at 60, 90 and 120 min; TNF-alpha was not detected in fluid from any of the animals that inhaled saline. As BAL fluids were toxic to the cells used in the assays, neither IL-1 nor IL-6 could be measured. We conclude that the monokine TNF-alpha is released into BAL fluid following anaphylactic challenge of passively sensitized guinea-pigs. The presence of the antiprotease, aprotinin, in the lavaging solution is essential for the detection and measurement of TNF-alpha in BAL fluid.  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