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 453 毫秒
1.
The two osteoclastogenesis pathways, receptor activator nuclear factor (NF)-kappaB ligand (RANKL)-mediated and fusion regulatory protein-1 (FRP-1)-mediated osteoclastogenesis, have recently been reported. There were significant differences in differentiation and activation mechanisms between the two pathways. When monocytes were cultured with FRP-1 without adding M-CSF, essential for the RANKL system, TRAP-positive polykaryocyte formation occurred. FRP-1-mediated osteoclasts formed larger pits on mineralized calcium phosphate plates than RANKL+M-CSF-mediated osteoclasts did. Lacunae on dentin surfaces induced by FRP-1-mediated osteoclasts were inclined to be single and isolated. However, osteoclasts induced by RANKL+M-CSF made many connected pits on dentin surfaces as if they crawled on there. Interestingly, FRP-1 osteoclastogenesis was enhanced by M-CSF/IL-1alpha, while chemotactic behavior to the dentin slices was not effected. There were differences in pH and concentration of HCO3- at culture endpoint and in adherent feature to dentin surfaces. Our findings indicate there are two types of osteoclasts with distinct properties.  相似文献   

2.
3.
4.
Because oxidative stress is a component of gastrointestinal injury, we investigated the effect of H(2)O(2) on transintestinal transport using isolated rat jejunum incubated in vitro. Millimolar concentrations of H(2)O(2) inhibited all the tested parameters without inducing any cytotoxic effect. Electrophysiological experiments indicated that H(2)O(2) decreases significantly both short circuit current and transepithelial electrical potential difference without affecting transepithelial resistance. The possibility that H(2)O(2) could influence (Na+, K+) -ATPase activity was explored using isolated basolateral membranes. Besides H(2)O(2), free radicals (O(2)(*-), HO*) were generated using different iron-dependent and independent systems; (Na+, K+) -ATPase activity was inhibited after membrane exposure to all ROS tested. The inhibition was prevented by allopurinol, superoxide dismutase or desferrioxamine. Western blot analysis showed a decreased expression of the alpha(1)-subunit of (Na+, K+) -ATPase. We conclude that H(2)O(2) may be a modulator of jejunal ion and water transport by multiple mechanisms, among which a significant inhibition of the basolateral (Na+, K+) -ATPase.  相似文献   

5.
To confirm whether or not the sulfo group of estradiol 17-sulfate (ES) is removed during in vivo metabolism in rats, the doubly labeled conjugate [6,7-3H, 35S] ES was injected into rats, and its biliary and urinary metabolites were determined by reverse isotope dilution method (RIDM). In male rats, the major radioactivity was detected in biliary disulfate fraction, which was composed of mainly ES and its two minor metabolites, 2-hydroxyestradiol 17-sulfate (2-OH-ES) and 2-methoxyestradiol 17-sulfate (2-MeO-ES). In female rats, in contrast, the radioactivity was dispersed into three fractions:biliary monosulfate, biliary disulfate, and urinary monosulfate fractions (Frs.) In both monosulfate Frs., 7beta-hydroxyestradiol 17-sulfate was detected as the major metabolite followed by 6alpha-, 6beta-, and 15beta-hydroxyestradiol 17-sulfates. Like male rats, 2-OH-ES and 2-Meo-ES as the minor products were detected in biliary disulfate fraction. The isotope ratios of ES and its metabolites in both sexes were essentially the same as that of the dose except that of 6alpha-hydroxylated metabolite, which may be derived from the loss of the tritium labeled at C6. These results confirm the occurrence of the direct metabolism of ES in rats.  相似文献   

6.
A full-length mosquito dopachrome conversion enzyme (DCE) and its truncated form lacking the last 54 carboxyl-terminal amino acid residues are expressed using a baculovirus/insect cell expression system. The full-length recombinant DCE displayed multiple bands during native PAGE with substrate staining, but only one active band was detected when the truncated recombinant DCE was analyzed under identical analysis conditions. Our data suggest that the last 50 some carboxyl-terminal residues are involved in the polymerization of the DCE molecules and that the proposed DCE isozymes likely reflect the presence of multimers of the same DCE molecules. The significance of the recombinant DCE in accelerating the melanization pathway is demonstrated by a rapid production of melanin in a dopa and tyrosinase reaction mixture in the presence of recombinant DCE. The DCE sequence data obtained in our previous study, together with results of functional expression and biochemical characterization achieved in this study, provide a necessary reference for the study of other insect DCEs.  相似文献   

7.
Nahar L  Turner AB 《Steroids》2003,68(14):1157-1161
Four lithocholic acid dimers were synthesised via esterification. The ester-linked dimer, 3-oxo-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta), was obtained by condensation of methyl lithocholate with 3-oxo-5beta-cholan-24-oic acid. Borohydride reduction of this ester-linked dimer gave 3alpha-hydroxy-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta), which was acetylated to 3alpha-acetoxy-5beta-cholan-24-oic acid (cholan-24-oic acid methyl ester)-3-yl ester, (3alpha,5beta). Reaction of methyl lithocholate with oxalyl chloride yielded the oxalate dimer, bis(5beta-cholan-24-oic acid methyl ester)-3alpha-yl oxalate.  相似文献   

8.
6,7-Dimethyl-8-ribityllumazine synthase (lumazine synthase) catalyses the penultimate step in the biosynthesis of riboflavin. In Bacillus subtilis, 60 lumazine synthase subunits form an icosahedral capsid enclosing a homotrimeric riboflavin synthase unit. The ribH gene specifying the lumazine synthase subunit can be expressed in high yield. All amino acid residues exposed at the surface of the active site cavity were modified by PCR assisted mutagenesis. Polar amino acid residues in direct contact with the enzyme substrates, 5-amino-6-ribitylamino-2,4(1H,3H)-pyrimidinedione and 3,4-dihydroxy-2-butanone 4-phosphate, could be replaced with relative impunity with regard to the catalytic properties. Only the replacement of Arg127, which forms a salt bridge with the phosphate group of 3,4-dihydroxy-2-butanone 4-phosphate, reduced the catalytic rate by more than one order of magnitude. Replacement of His88, which is believed to assist in proton transfer reactions, reduced the catalytic activity by about one order of magnitude. Surprisingly, the activation enthalpy deltaH of the lumazine synthase reaction exceeds that of the uncatalysed reaction. On the other hand, the free energy of activation deltaG of the uncatalysed reaction is characterised by a large entropic term (TdeltaS) of -37.8 kJmol(-1), whereas the entropy of activation (TdeltaS) of the enzyme-catalysed reaction is -6.7 kJmol(-1). This suggests that the rate enhancement by the enzyme is predominantly achieved by establishing a favourable topological relation of the two substrates, whereas acid/base catalysis may play a secondary role.  相似文献   

9.
This study shows the presence of all three nitric oxide synthases (NOSs) and NOS activity in H9c2 cells cultured under non-stimulated conditions. By using the 4,5 diaminofluoresceindiacetate (DAF-2DA) fluorimetric nitric oxide (NO(*)) detection system we observed NO(*) production in H9c2 cells. As revealed by confocal microscopy, NO(*) fluorescence colocalizes in mitochondria labeled with Mito-Tracker Red CM-H(2)Xros. Upon stimulation with acetylcholine (Ach), which increased NOS activity by 75%, the colocalization coefficient C(green) value, calculated as Pearson's correlation, increased from 0.07 to 0.10, demonstrating an augmented presence of NO(*) in mitochondria. Conversely, the presence of NO(*) in mitochondria decreased following cells pretreatment with l-MonoMethylArginine (L-NMMA), a competitive inhibitor of NOS activity, as indicated by the reduction of the C(green) value to 0.02. This work confirms that the presence of NO(*) in mitochondria can be modulated in response to different fluxes of NO(*).  相似文献   

10.
Cells of the coccolithophorid Emiliania huxleyi strain CS-57 grown under an atmosphere of air+0.5% CO(2) showed oxidative damage after 10 days growth with concomitant and major changes to the lipid composition. The fatty acid profile was strongly altered and lacked appreciable amounts of the polyunsaturated fatty acids (PUFA: C(18:5), C(18:3) and C(22:6)) typical of healthy cells. Oxidation products of these PUFA could not be detected, but monounsaturated fatty acids proved to be good indicators of oxidative processes. The presence (after NaBH(4)-reduction) of a high proportion of 11-hydroxyoctadec-cis-9-enoic and 8-hydroxyoctadec-cis-9-enoic acids showed that the degradation of oleic acid involved mainly free radical oxidation processes (70-75% autoxidation and 20-25% photooxidation). We also detected large amounts of degradation products of the oxidation product 9,10-epoxyoctadecanoic acid including diols, methoxyhydrins and chlorohydrins. These oxidative effects were found in all the lipid classes examined. Products included significant amounts of chlorophyll side-chain autooxidation products Z- and E-3,7,11,15-tetramethylhexadec-3-ene-1,2-diols and Z-and E-3,7,11,15-tetramethylhexadec-2-ene-1,4-diols, while phytyldiol was present in relatively low proportions. Delta(5)-3beta,7-epimeric unsaturated steroidal diols arising from the autooxidation of the Delta(5) double bond of epi-brassicasterol and minor amounts of Delta(4)-3beta,6-diols were also detected. Long-chain unsaturated ketone (alkenone) content per cell was much higher in the presence of 0.5% CO(2) likely due to carbon storage under these conditions. The proportions of di- and tri-unsaturated alkenones was relatively stable throughout the growth cycle in the absence of additional CO(2), but not when grown with 0.5% CO(2). The detection of characteristic alkenone autoxidation products in cells grown under these latter conditions allowed us to attribute the significant increase in index observed to the involvement of free radical oxidation processes.  相似文献   

11.
Pressure-induced unfolding of 23-kDa protein from spinach photosystem II has been systematically investigated at various experimental conditions. Thermodynamic equilibrium studies indicate that the protein is very sensitive to pressure. At 20 degrees C and pH 5.5, 23-kDa protein shows a reversible two-state unfolding transition under pressure with a midpoint near 160 MPa, which is much lower than most natural proteins studied to date. The free energy (DeltaG(u)) and volume change (DeltaV(u)) for the unfolding are 5.9 kcal/mol and -160 ml/mol, respectively. It was found that NaCl and sucrose significantly stabilize the protein from unfolding and the stabilization is associated not only with an increase in DeltaG(u) but also with a decrease in DeltaV(u). The pressure-jump studies of 23-kDa protein reveal a negative activation volume for unfolding (-66.2 ml/mol) and a positive activation volume for refolding (84.1 ml/mol), indicating that, in terms of system volume, the protein transition state lies between the folded and unfolded states. Examination of the temperature effect on the unfolding kinetics indicates that the thermal expansibility of the transition state and the unfolded state of 23-kDa protein are closer to each other and they are larger than that of the native state. The diverse pressure-refolding pathways of 23-kDa protein in some conditions were revealed in pressure-jump kinetics.  相似文献   

12.
As typical anchorage-dependent cells myocytes must balance contractility against adequate adhesion. Skeletal myotubes grown as isolated strips from myoblasts on micropatterned glass exhibited spontaneous peeling after one end of the myotube was mechanically detached. Such results indicate the development of a prestress in the cells. To assess this prestress and study the dynamic adhesion strength of single myocytes, the shear stress of fluid aspirated into a large-bore micropipette was then used to forcibly peel myotubes. The velocity at which cells peeled from the surface, V(peel), was measured as a continuously increasing function of the imposed tension, T(peel), which ranges from approximately 0 to 50 nN/ micro m. For each cell, peeling proved highly heterogeneous, with V(peel) fluctuating between 0 micro m/s ( approximately 80% of time) and approximately 10 micro m/s. Parallel studies of smooth muscle cells expressing GFP-paxillin also exhibited a discontinuous peeling in which focal adhesions fractured above sites of strong attachment (when pressure peeled using a small-bore pipette). The peeling approaches described here lend insight into the contractile-adhesion balance and can be used to study the real-time dynamics of stressed adhesions through both physical detection and the use of GFP markers; the methods should prove useful in comparing normal versus dystrophic muscle cells.  相似文献   

13.
Activation of the O(2)(-) generating NADPH oxidase of phagocytes results from the assembly of the membrane-bound flavocytochrome b(558) with cytosolic proteins, p67(phox), p47(phox), and Rac. However, it has been recently reported that the arachidonic acid- and calcium-binding heterodimer S100A8/A9, abundant in neutrophil cytosol, influences the activation process. In a semi-recombinant system comprising neutrophil membranes, recombinant proteins, p67(phox), p47(phox), GTPgamma S-loaded Rac2, and arachidonic acid (AA), both the rate and the extent of the oxidase activation were increased by S100A8/A9, provided it was preloaded with AA. Binding of [(14)C]AA to S100A8/A9 was potentiated by recombinant cytosolic phox proteins and GTPgammaS, suggesting the formation of a complex, comprising oxidase activating proteins and S100A8/A9, with a greater affinity for AA. The rate constant of oxidase activation was not increased by AA-loaded S100A8/A9, whereas the maximal oxidase activity elicited was twice as high. AA-loaded S100A8/A9 increases oxidase activation probably by decreasing the deactivation rate.  相似文献   

14.
Patterns of salivary HCO(3)(-) secretion vary and depend on species and gland types. However, the identities of the transporters involved in HCO(3)(-) transport and the underlying mechanism of intracellular pH (pH(i)) regulation in salivary glands still remain unclear. In this study, we examined the expression of the Na(+)-HCO(3)(-) cotransporter (NBC) and its role in pH(i) regulation in guinea pig salivary glands, which can serve as an experimental model to study HCO(3)(-) transport in human salivary glands. RT-PCR, immunohistochemistry, and pH(i) measurements from BCECF-AM-loaded cells were performed. The amiloride-sensitive Na(+)/H(+) exchanger (NHE) played a putative role in pH(i) regulation in salivary acinar cells and also appeared to be involved in regulation in salivary ducts. In addition to NHE, NBC also played a role in pH(i) regulation in both acini and ducts. In the parotid gland, NBC1 was functionally expressed in the basolateral membrane (BLM) of acinar cells and the luminal membrane (LM) of ducts. In the submandibular gland, NBC1 was expressed only in the BLM of ducts. NBC1 expressed in these two types of salivary glands takes up HCO(3)(-) and is involved in pH(i) regulation. Although NBC3 immunoreactivity was also detected in submandibular gland acinar cells and in the ducts of both glands, it is unlikely that NBC3 plays any role in pH(i) regulation. We conclude that NBC1 is functionally expressed and plays a role in pH(i) regulation in guinea pig salivary glands but that its localization and role are different depending on the type of salivary glands.  相似文献   

15.
It is generally believed that compatible osmolytes stabilize proteins by shifting the denaturation equilibrium, native state <--> denatured state toward the left. We show here that if osmolytes are compatible with the functional activity of the protein at a given pH and temperature, they should not significantly perturb this denaturation equilibrium under the same experimental conditions. This conclusion was reached from the measurements of the activity parameters (K(m) and k(cat)) and guanidinium chloride-induced denaturations of lysozyme and ribonuclease-A in the presence of five polyols (sorbitol, glycerol, mannitol, xylitol and adonitol) at pH 7.0 and 25 degrees C.  相似文献   

16.
Effects of cellulase on the modification of cellulose   总被引:1,自引:0,他引:1  
Cao Y  Tan H 《Carbohydrate research》2002,337(14):1291-1296
Multicomponent cellulases, purified endoglucanases and cellobiohydrolases were assayed and shown to modify pure natural cellulose (softwood pulp). Changes in structure and properties of the cellulose caused by enzymatic treatment depend on the composition, the type of enzyme, and the treatment conditions. The reactivity of cellulose for some dissolving and derivatization processes may be improved by enzymatic hydrolysis. Endoglucanases decreased the average degrees of polymerization (DP) and improved the alkaline solubility of cellulose most efficiently. The variation in the supramolecular structure estimated from the infrared spectra of the cellulose samples was found to be correlated with the reactivity and might represent wide variations in conformation caused by the breakdown of the hydrogen bonds.  相似文献   

17.
Bollo M  Bonansea S  Machado EE 《FEBS letters》2006,580(11):2686-2690
We studied the effect of Na(+) extracellular on Ca(2+) mobilization from intracellular store evoked by carbachol in Trypanosoma cruzi. We report that slow component of Ca(2+) signaling evoked by agonist is dependent on extracellular Na(+) but not on InsP(3) increase. Moreover, this Ca(2+) signaling progressively increased when pH of the medium changed from 7.0 to 7.8. In addition, we found that it was regulated by PKC. The agonist was also able to induce the alkalinization of the acidic compartment, and both Ca(2+) signaling and alkalinization were inhibited by the EIPA-inhibitor of the Na(+)/H(+) exchanger. These results demonstrated the alkalinization of acidic vacuoles and PKC are involved in the triggering of the epimastigote Ca(2+) signaling.  相似文献   

18.
This study showed that four factors which stimulate transepithelial fluid secretion and inorganic ion transport across the main segment of the Malpighian tubules of Drosophila melanogaster also stimulate transepithelial secretion of the prototypical organic cation tetraethylammonium (TEA). TEA fluxes across the Malpighian tubules and gut were measured using a TEA-selective self-referencing (TEA-SeR) microelectrode. TEA flux across isolated Malpighian tubules was also measured using a TEA-selective microelectrode positioned in droplets of fluid secreted by tubules set up in a modified Ramsay assay. TEA flux was stimulated by the intracellular second messengers cAMP and cGMP, which increase the lumen-positive transepithelial potential (TEP), and also by tyramine and leucokinin-I (LK-I), which decrease TEP. The largest increase was measured in response to 1 micromol l-1 LK-I which increased transepithelial TEA flux by 72%. TEA flux in the lower tubule was stimulated slightly (13%) by 1 micromol l-1 tyramine but not by any of the other factors. TEA flux across the midgut was unaffected by cAMP, cGMP or tyramine. This is the first study to demonstrate the effects of insect diuretic factors and second messengers on excretion of organic cations.  相似文献   

19.
Ion transporters such as Na(+)/H(+) exchanger (NHE), Cl(-)/HCO(3)(-) exchanger (AE), and Na(+)/HCO(3)(-) cotransporter (NBC) are known to contribute to the intracellular pH (pH(i)) regulation during agonist-induced stimulation. This study examined the mechanisms for the pH(i) regulation in the mouse parotid and sublingual acinar cells using the fluorescent pH-sensitive probe, BCECF. The pH(i) recovery from agonist-induced acidification in the sublingual acinar cells was completely blocked by EIPA, a NHE inhibitor. However, the parotid acinar cells required DIDS, a NBC1 inhibitor, in addition to EIPA in order to block the pH(i) recovery. Moreover, RT-PCR analysis detected the expression of pancreatic NBC1 (pNBC1) only in the parotid acinar cells. These results provide strong evidence that the mechanisms for the pH(i) regulation are different in the two types of acinar cells, and pNBC1 contributes to pH(i) regulation in the parotid acinar cells, whereas NHE is likely to be the exclusive pH(i) regulator in the sublingual acinar cells.  相似文献   

20.
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