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 46 毫秒
1.
The (Na++K+)-activated, Mg2+-dependent ATPase from rabbit kidney outer medulla was prepared in a partially inactivated, soluble from depleted of endogenous phospholipids, using deoxycholate. This preparation was reactivated 10 to 50-fold by sonicated liposomes of phosphatidylserine, but not by non-sonicated phosphatidylserine liposomes or sonicated phosphatidylcholine liposomes. The reconstituted enzyme resembled native membrane preparations of (Na++K+)-ATPase in its pH optimum being around 7.0 showing optimal activity at Mg2+: ATP mol ratios of approximately 1 and a Km value for ATP of 0.4 mM.Arrhenius plots of this reactivated activity at a constant pH of 7.0 and an Mg2+: ATP mol ratio of 1:1 showed a discontinuity (sharp change of slope) at 17 °C, With activation energy (Ea) values of 13–15 kcal/mol above this temperature and 30–35 kcal below it. A further discontinuity was also found at 8.0 °C and the Ea below this was very high (> 100 kcal/mol).Incresed Mg2+ concentrations at Mg2+: ATP ratios in excess of 1:1 inhibited the (Na++K+)-ATPase activity and also abolished the discontinuities in the Arrhenius plots.The addition of cholesterol to phosphatidylserine at a 1:1 mol ratio partially inhibited (Na++K+)-ATPase reactivation. Arrhenius plots under these conditions showed a single discontinuity at 20°C and Ea values of 22 and 68kcal/mol above and below this temperature respectively. The ouabain-insensitive Mg2+-ATPase normally showed a linear Arrhenius plot with an Ea of 8 kcal/mol. The cholesterol-phosphatidylserine mixed liposomes stimulated the Mg2+-ATPase activity, which now also showed a discontinuity at 20 °C with, however, an increased value of 14 kcal/mol above this temperature and 6 kcal/mol below. Kinetic studies showed that cholesterol had no significant effect on the Km for ATP.Since both of cholesterol and Mg2+ are know to alter the effects of temperature on the fluidity of phospholipids the above result are discussed in this context.  相似文献   

2.
Basic substances and acidic ionophores that increase the lysosomal pH in cultured macrophages (Ohkuma, S., and B. Poole, 1978, Proc. Natl. Acad. Sci. USA., 75:3327-3331; Poole, B., and S. Ohkuma, 1981, J. Cell Biol., 90:665-669) inhibited the digestion of heat-denatured acetylated bovine serum albumin (BSA) taken up by the cells. For several substances, the shift in pH sufficed to explain the inhibition of proteolysis. Additional effects, presumably on enzyme activities, have to be postulated for tributylamine, amantadine, and chloroquine. Sodium fluoride (10 mM) had no significant effect on the breakdown of BSA by macrophages. The breakdown of endogenous macrophage proteins, whether short lived or long lived, was inhibited approximately 40% by 10 mM NaF and 30%, or sometimes less in the case of long-lived proteins, by 100 microM chloroquine. When the cells were supplied with BSA, a mixture of cell proteins, or even inert endocytosible materials, the breakdown of endogenous long-lived proteins and the inhibitory effect of chloroquine on this process were selectively reduced. Inhibition of endocytosis by cytochalasins B or D did not affect the chloroquine-sensitive breakdown of endogenous proteins, indicating that the proteins degraded by this process were truly endogenous and not taken in from the outside by cellular cannibalism. On the other hand, when macrophage proteins were supplied extracellularly, their breakdown occurred at the same rate for short-lived and long-lived proteins, and it was strongly inhibited by chloroquine and not by NaF. It is concluded from these results that the breakdown of endogenous proteins, both short-lived and long-lived, probably takes place partly (approximately 30%) in lysosomes and partly through one or more nonlysosomal mechanism(s) unaffected by chloroquine and presumably susceptible to inhibition by fluoride. A difference must exist between short-lived and long-lived proteins in the manner in which they reach lysosomes or are handled by these organelles; this difference would account for the selective effect of the supply of endocytosible materials on the lysosomal processing of long-lived proteins.  相似文献   

3.
The subunit MW of Dioscorea bulbifera polyphenol oxidase (MW 115 000 ± 2000) determined by SDS-PAGE is ca. 31 000 indicating that the enzyme is an oligomeric protein with four subunits. Ki values of various inhibitors and their modes of inhibition have been determined with catechol and pyrogallol as substrates. p-Nitrophenol, p-cresol, quinoline and resorcinol are competitive inhibitors of catechol binding while only orcinol and p-nitrophenol behave in the same way towards pyrogallol as substrate. From the effect of pH on Vmax, groups with pK values ca. 4.7 and 6.8 have been identified to be involved in catalytic activity. The Arrhenius activation energy (Ea) at pH 4.0 is 8.9 kcal/mol between 40–65°. At pH 7.0, the value is 22.1 kcal/mol between 40 and 60°. The enthalpies (ΔH) at pH 4.0 and pH 7.0 are 2.3 kcal/mol and 32.4 kcal/mol respectively. The results are discussed considering the conformational changes of the enzyme during substrate binding.  相似文献   

4.
Isotopic exchange kinetics at equilibrium for E. coli native aspartate transcarbamylase at pH 7.8, 30 °C, are consistent with an ordered BiBi substrate binding mechanism. Carbamyl phosphate binds before l-Asp, and carbamyl-aspartate is released before inorganic phosphate. The rate of [14C]Asp C-Asp exchange is much faster than [32P]carbamyl phosphate Pi exchange. Phosphate, and perhaps carbamyl phosphate, appears to bind at a separate modifier site and prevent dissociation of active-site bound Pi or carbamyl phosphate. Initial velocity studies in the range of 0–40 °C reveal a biphasic Arrhenius plot for native enzyme: Ea (>15 °C) = 6.3 kcal/ mole and Ea (<15 °C) = 22.1 kcal/mole. Catalytic subunits show a monophasic plot with Ea ? 20.2 kcal/mole. This, with other data, suggests that with native enzyme a conformational change accompanying aspartate association contributes significantly to rate limitation at t > 15 °C, but that catalytic steps become definitively slower below 15 °C. Model kinetics are derived to show that this change in mechanism at low temperature can force an ordered substrate binding system to produce exchange-rate patterns consistent with a random binding system with all exchange rates equal. The nonlinear Arrhenius plot also has important consequences for current theories of catalytic and regulatory mechanisms for this enzyme.  相似文献   

5.
The thermotropic properties of bovine blood coagulation Factors IX and X, as well as the activation intermediates and products of these proteins, have been investigated by differential scanning microcalorimetry in the presence and absence of Ca2+. Bovine Factor IX displays a single thermal-denaturation transition characterized by a temperature midpoint (TM) of 54.5 ± 0.5 °C and a calorimetric enthalpy (ΔHc) of 105 ± 15 kcal/mol, in the absence of Ca2+. In the presence of Ca2+ concentrations sufficient to saturate its sites on Factor IX, the Tm value is increased to 57.0 ± 0.5 °C and the ΔHc is virtually unchanged. When the activation intermediate, Factor IXα, is similarly analyzed in the absence of Ca2+, a broad, diffuse thermogram was obtained which did not lend itself to calculation of thermodynamic parameters. In the presence of Ca2+, Factor IXα displayed thermograms characterized by a TM of 51.0 ± 0.5 °C and a ΔHc of 109 ± 10 kcal/mol. The activated product, Factor IXaα, in the absence of Ca2+ (the values in the presence of saturating Ca2+ are given in parentheses), undergoes thermal denaturation with a TM of 54.5 ± 0.5 °C (57.0 ± 0.5 °C) and a ΔHc of 158 ±10 kcal/mol (156 ± 10 kcal/mol). Similarly, the terminal-activation product, Factor IXaβ, displays a TM of 51.5 ± 0.5 °C (54.0 ± 0.5 °C) and a ΔHc of 85 ± 5 kcal/mol (126 ± 10 kcal/mol). Bovine blood coagulation Factor X has been analyzed in this same fashion, and shows very similar thermal properties to Factor IX. The thermal denaturation of Factor X is represented by a TM of 54.0 ± 0.5 °C (55.0 ± 0.5 °C) and a ΔHc of 102 ± 10 kcal/mol (118 ± 10 kcal/mol), whereas its activated form, Factor Xaβ, possesses a TM of 55.0 ± 0.5 °C (55.0 ± 0.5 °C) and a ΔHc of 92.0 ± 5 kcal/mol (136 ± 10 kcal/mol). These studies indicate that, for many of these proteins, Ca2+ induces a conformational alteration to a more thermally stable form, which also requires the absorption of greater amounts of heat for thermal denaturation.  相似文献   

6.
The rate of ouabain-resistant Li+-efflux was studied in erythrocytes of normal controls and of patients with essential hypertension. Despite variability in rate, erythrocytes from normotensive persons revealed a uniform pattern of temperature dependence of the efflux, with two slopes (Ka = 9.4 and 19.1 kcal/mol, respectively) and a transition at about 25°C. Erythrocytes from the patients showed both a higher rate of Li+ efflux and significant changes in the temperature repsonse, with essentially a single slope (Ka = 14 kcal/mol). The data indicate localized changes in the membrane organization of hypertensive erythrocytes, involving lipid-protein interaction.  相似文献   

7.
The fluorescent dye rhodamine 123 (R123) decreases the intracellular ATP levels and also inhibits the degradation of short-lived proteins in isolated hepatocytes. This inhibition affects lysosomal and, to some extent, non-lysosomal mechanisms. The degradation of short-lived proteins decreases more when ATP levels are less than 40% of those in control cells, in contrast to the reported linear correlation between ATP levels and degradation of long-lived proteins. R123 provides a powerful probe for clarifying the proteolytic mechanisms involved in degradation of short-lived proteins and the ATP requirements in protein degradation. Indeed, as illustrated, the results suggest different mechanisms for the degradation of short- and long-lived proteins. Moreover, they provide a warning for the clinical use of this reagent.  相似文献   

8.
Low density lipoprotein (LDL) processing has been investigated in the subcloned human colonic carcinoma cell line HT29-18. LDL binding at 4°C was a saturable process in relation to time and LDL concentration. The Kd for LDL binding was 11 g/ml. ApoE-free HDL3 or acetylated LDL did not significantly compete with125I-LDL binding, up to 500 g/ml.125I-LDL binding was decreased by 70% in HT29-18 cells preincubated for 24 hours in culture medium containing 100 g/ml unlabelled LDL. Ligand blotting studies performed on HT29-18 homogenates using colloidal gold labelled LDL indicated the presence of one autoradiographic band corresponding to an apparent molecular weight of 130 kDa, which is consistent with the previously reported molecular weight of the LDL receptor in human fibroblasts. At 37°C,125I-LDL was actively internalized by HT29-18 cells and lysosomal degradation occurred as demonstrated by the inhibitory effect of chloroquine. LDL uptake and degradation by HT29-18 cells also resulted in a marked decrease in endogenous sterol synthesis. These data demonstrate that the HT29-18 human cancerous intestinal cells are able to specifically bind and internalize LDL, and that LDL processing results in down-regulation of sterol biosynthesis. Thus, intestinal epithelial cells possess specific LDL receptors that can be exploited to accomplish drug delivery and gene transfer via the receptor-mediated endocytosis pathway.Abbreviations HDL, HCL3 high density lipoprotein - LDL low density lipoprotein  相似文献   

9.
Mollusc shells represent excellent systems for the preservation and retrieval of genuine biomolecules from archaeological or palaeontological samples. As a consequence, the post-mortem breakdown of intracrystalline mollusc shell proteins has been extensively investigated, particularly with regard to its potential use as a “molecular clock” for geochronological applications. But despite seventy years of ancient protein research, the fundamental aspects of diagenesis-induced changes to protein structures and sequences remain elusive. In this study we investigate the degradation of intracrystalline proteins by performing artificial degradation experiments on the shell of the thorny oyster, Spondylus gaederopus, which is particularly important for archaeological research. We used immunochemistry and tandem mass tag (TMT) quantitative proteomics to simultaneously track patterns of structural loss and of peptide bond hydrolysis.Powdered and bleached shell samples were heated in water at four different temperatures (80, 95, 110, 140 °C) for different time durations. The structural loss of carbohydrate and protein groups was investigated by immunochemical techniques (ELLA and ELISA) and peptide bond hydrolysis was studied by tracking the changes in protein/peptide relative abundances over time using TMT quantitative proteomics. We find that heating does not induce instant organic matrix decay, but first facilitates the uncoiling of cross-linked structures, thus improving matrix detection. We calculated apparent activation energies of structural loss: Ea (carbohydrate groups) = 104.7 kJ/mol, Ea (protein epitopes) = 104.4 kJ/mol, which suggests that secondary matrix structure degradation may proceed simultaneously with protein hydrolysis. While prolonged heating at 110 °C (10 days) results in complete loss of the structural signal, surviving peptide sequences were still observed. Eight hydrolysis-prone peptide bonds were identified in the top scoring shell sequence, the uncharacterised protein LOC117318053 from Pecten maximus. Interestingly, these were not the expected “weak” bonds based on published theoretical stabilities calculated for peptides in solution. This further confirms that intracrystalline protein degradation patterns are complex and that the overall microchemical environment plays an active role in protein stability. Our TMT approach represents a major stepping stone towards developing a model for studying protein diagenesis in biomineralised systems.  相似文献   

10.
In this work different aspects of the glucose-fructose enzymatic isomerization, using immobilized glucose isomerase, are studied and quantified. Reaction temperatures range from 40?°C to 60?°C. Intra-particle effective diffusivities (D e), determined after uptake experiments, are between 1.20?×?10?6?cm2/s, at 40?°C, and 2.52?×?10?6?cm2/s, at 60?°C. The estimated energy of activation for diffusion (E aD) is 7.71?kcal/mol. No significant adsorption of the sugars on the support gel matrix is observed. Crushed particles (φ = 150–350?μ) are used during kinetic experiments. For this range of particle diameters, inherent kinetics is approached. A reversible Michaelis–Menten rate equation is fitted to the data, providing the following parameters at pH = 7.0: k 0 = 2.15?×?10?6?g/IU/s; E a/R = 8998?K. Glucose (K G) and fructose (K F) affinity constants are essentially the same, ranging from 0.190?M, at 40?°C to 0.305?M, at 60?°C. The thermodynamic equilibrium constant is determined for the three temperatures, and the heat of reaction, estimated from a Van't Hoff plot, is ΔH = 1682?cal/mol. Independent experiments, where the reaction occurs in the presence of significant intra-particle mass transfer resistance, are used as validation tests.  相似文献   

11.
(LDH) obtained from larvae of Drosophila melanogaster was purified to homogeneity by affinity chromatography on oxamate-Sepharose. The purification procedure is simple to operate and gives a homogeneous preparation in a good yield (34.86%) after only two steps. Utilizing the homogeneous LDH preparation, an attempt was made to characterize the LDH molecule. Thus, it was found that the N-terminal amino acid is isoleucine, and the enzyme is tetrameric and composed of four identical subunits of apparent molecular weight 38,000, suggesting that it is controlled by a single gene. Homogeneous LDH preparations exhibit one band on neutral acrylamide gels when the substrate is either dl-lactic acid or l-(+)-lactate. The optimum temperature is 45°C for the purified enzyme and 40°C for the crude homogenate. The K m values for pyruvate and NADH are 0.154 and 0.027mm, respectively, while the K m values for lactate and NAD are 29.4 and 1.33mm, respectively. A discontinuity in the E a slope was observed at a transition temperature of 30°C. The E a value between 20 and 30°C was calculated as 12.06 kcal/mol, while between 30 and 45°C the E a value was 4.01 kcal/mol. This evidence, together with other observations reported in the literature, suggests that the LDHs of invertebrates and vertebrates have arisen by divergent evolution from a common ancestral gene.  相似文献   

12.
Coupling the cathepsin D inhibitor pepstatin to poly(l-Lys) (Mr 13,000) is shown to enhance its inhibition of protein breakdown in whole cell systems. Rates of intracellular protein breakdown for prelabeled proteins of Balb/c 3T3 fibroblasts were measured in the presence and absence of amino acids and insulin to generate basal and enhanced rates of protein breakdown. Pepstatin and poly(l-Lys) inhibited rates of degradation 5–7% and 16–23%, respectively, under each condition. Pepstatinyl-poly(l-Lys), containing 9 mol pepstatin/mol polymer, inhibited enhanced rates of degradation a further 24–33% compared to poly(l-Lys), but this extra increment was not seen under basal conditions. Although the mechanism of inhibition of intracellular protein breakdown by poly(l-Lys) presently is unknown, the data obtained with free and conjugated pepstatin indicate the lysosomal system degrades proteins under both basal and enhanced conditions.  相似文献   

13.
A GH 26 endo-mannanase from Bacillus sp. CFR1601 was purified to homogeneity (Mw ∼39 kDa, specific activity 10,461.5 ± 100 IU/mg). Endo-mannanase gene (manb-1601, 1083 bp, accession No. KM404299) was expressed in Escherichia coli BL21 (DE3) and showed typical fingerprints of α/β proteins in the far-UV CD. A high degree of conservation among amino acid residues involved in metal chelation (His-1, 23 and Glu-336) and internal repeats (122–152 and 181–212) was observed in endo-mannanases reported from various Bacillus sp. Thermal inactivation kinetics suggested that metal ions are quintessential for stabilization of ManB-1601 structure as holoenzyme (Ea 87.4 kcal/mol, ΔH 86.7 kcal/mol, ΔS 186.6 cal/k/mol) displayed better values of thermodynamic parameters compared to metal-depleted ManB-1601 (Ea 47 kcal/mol, ΔH 45.7 kcal/mol, ΔS 64.7 cal/k/mol). EDTA treatment of ManB-1601 not only lead to transitions in both secondary and tertiary structure but also promulgated the population of conformational state that unfolds at lower temperature. ManB-1601 followed a three-state process for thermal inactivation wherein loss of tertiary structure preceded the concurrent loss of secondary structure and activity.  相似文献   

14.
Rat promegakaryoblasts (RPM, a precursor platelet cell line) in culture exhibited a capacity to bind, take up and degrade125I-LDL. The low density lipoprotein (LDL) binding showed the following characteristics: (a) high affinity, (b) saturability, (c) specificity, (d) down-regulation, after exposure to 25 hydroxycholesterol. Furthermore the proteolytic degradation of125I-LDL by RPMs was inhibited by chloroquine which interferes with the lysosomal degradation processes. These findings show LDL receptor cell biology of RPM to be of the classical type and to differ from that of platelets.  相似文献   

15.
《Bioorganic chemistry》1987,15(2):100-108
Nonenzymatic rates of hydrolytic deamination of adenosine and cytidine by acids and bases analogous to side chains of naturally occurring amino acids are compared with the rates of uncatalyzed deamination in water and with the rates of the hydroxide- and hydrogen ion-catalyzed reactions. For adenosine, hydroxide ion is an effective catalyst, with a second-order rate constant of 7.5 × 10−6 m−1 s−1 at 85°C and an energy of activation of 19.9 kcal/mol. Acid-catalyzed deamination of adenine proceeds with a second-order rate constant of 1.5 × 10−6 m−1 s−1 at 85°C. At concentrations of 1 m and at pH values corresponding to their respective pKa values, dimethylamine, acetate, selenide, imidazole, phosphate, and zinc(II) do not enhance the rate of deamination of adenosine beyond that observed in water, and 2-mercaptoethanol produces only a modest rate enhancement. The uncatalyzed rate of adenosine deamination in water is 8.6 × 10−9 s−1 at 85°C: extrapolation to 37°C and comparison with kcat for rat hepatoma adenosine deaminase yield a rate enhancement by the enzyme of approximately 2 × 1012-fold. 1,6-Dimethyladenosine, the conjugate acid of which has a pKa value much higher than that of adenosine, is not readily deaminated, suggesting that the uncatalyzed deamination of adenosine does not proceed by hydroxide ion attack on the rare protonated form of adenosine, but rather by attack on the neutral species. Deamination of cytidine is catalyzed most effectively by hydroxide ion, with a second-order rate constant of 4.5 × 10−4 m−1 s−1 at 85°C and an energy of activation of 28.5 kcal/mol. The uncatalyzed rate of deamination of cytidine in water, which also exhibits an energy of activation of 28.5 kcal/mol, is 8.8 × 10−8 s−1 at 85°C. Comparison of the rate extrapolated to 25°C with kcat for bacterial cytidine deaminase gives a rate enhancement for the enzyme of 4 × 1011-fold. The C-5 proton of the pyrimidine ring of cytidine does not exchange with solvent during alkaline hydrolysis, suggesting that deamination under these conditions does not involve prior addition of water across the 5,6 double bond.  相似文献   

16.
The trinuclear clusters of general composition [Ru3O(OOCCH3)6(N-Het)3], where N-Het=pyridine and pyrazine derivatives, exhibit a series of reversible waves in the range of −1.8 to 2.4 V versus SHE, in acetonitrile, ascribed to the successive [cluster]−2/−1/0/+1/+2/+3 redox couples. The redox potentials decrease with the pKa of the N-heterocyclic ligands according to the equations E°(+3/+2)= 2.24−0.023 pKa; E°(+2/+1)=1.34−0.029 pKa; E°(+1/0)=0.36−0.039 pKa and E°(0/−1)=−0.68− 0.074 pKa. The dependence is greater at lower oxidation states, reflecting the role of π-backbonding in the complexes.  相似文献   

17.
The occurrence of a receptor for human LDL was investigated in the tegument of adult Schistosoma mansoni employing several approaches. Binding of LDL to SDS-PAGE fractionated tegument proteins was measured directly on nitro-cellulose membranes and visualised by an anti-human LDL antibody. Proteins with an Mr of 60, 35 and 14 kDa were evidenced. Affinity chromatography of 125 I-labelled tegument proteins on a LDL-Sepharose column, revealed the same pattern of proteins observed in the immunoblot experiments. Finally, the binding of human LDL to the intact tegument was measured by microcalorimetry. Binding was shown to be an exothermic reaction, releasing approximately 2500 kcal/mol, it was saturable, and reproducibly displayed a biphasic curve suggesting that binding of LDL to S. mansoni might occur through a two step process, initiated by a nonspecific hydrophobic interaction followed by a specific high affinity ligand-receptor reaction. Pre-treatment of the tegument with trypsin reduced the binding of LDL to the tegument. Furthermore, albumin, which is an abundant lipid carrier protein in the serum and thus a potential ligand, failed to release any measurable heat when incubated with the tegument.  相似文献   

18.
Pseudomonas putida (MTCC 1194) has been used to degrade phenol in water in the concentration range 100–1000?ppm. The inhibition effects of phenol as substrate have become predominant above the concentration of 500?ppm (5.31?mmoles/dm3). The optimum temperature and initial pH required for maximum phenol biodegradation were 30?°C and 7.00 respectively. From the degradation data the activation energy (E a ) was found to be equal to 13.8?kcal/g mole substrate reacted. The most suitable inoculum age and volume for highest phenol degradation were 12?hrs and 7% v/v respectively. Surfactants had negligible effect on phenol biodegradation process for this microorganism. Monod model has been used to interpret the free cell data on phenol biodegradation. The kinetic parameters have been estimated upto initial concentration of 5.31?mmoles/dm3. μ max and K S gradually increased with higher concentration of phenol. However, beyond the phenol concentration of 5.31?mmoles/dm3, the inhibition became prominant. The μ max has been to be a strong function of initial phenol concentration. The simulated and the experimental phenol degradation profiles have good correspondence with each other.  相似文献   

19.
ABSTRACT

Chondroitin sulfate (CS) used for treatment of osteoarthritis exerts distinct effects on human articular chondrocytes in vitro. We performed a binding analysis with 99mTc-labeled CS (Condrosulf, a commercial CS preparation containing calcium stearate) and cultured human chondrocytes in order to evaluate the presence of specific receptors. Saturation binding at 37°C for 2?h revealed the presence of high-affinity binding sites for CS with a Kd of 2.3 × 10?9?mol/L and a Bmax of 5.0 × 108. Extensive dialysis of Chondrosulf led to a decrease of the binding affinity by 52.5 ± 19.5% and of the number of CS binding sites/cell by 62.0 ± 14.0%, demonstrating that the additive present in the Condrosulf preparation enhances CS binding. The nature of the binding site is not yet known but evidence exists in the literature that the scavenger receptor CD36, thoroughly investigated on macrophages, is also found on chondrocytes and might be involved in CS binding. Therefore, we undertook a comparative binding study with human monocytes and labelled LDL and oxidized LDL, the latter being a postulated atherogenic agent in atherosclerosis. For [125I]-LDL binding we found a Kd of 0.45 × 10?8?mol/L and a Bmax of 0.14 × 106 on quiescent monocytes and for [125I]-(ox)LDL binding a Kd of 1.8 × 10?8?mol/L and a Bmax of 1.3 × 106 using LPS-activated monocytes. These data are comparable to the binding affinity found for lipoprotein–proteoglycan-complexes and hence are an indication but not a proof that CD36 is involved in CS binding to human chondrocytes.  相似文献   

20.
Degradation of proteins in the cells occurs by proteasomes, lysosomes and other cytosolic and organellar proteases. It is believed that proteasomes constitute the major proteolytic pathway under most conditions, especially when degrading abnormal and other short-lived proteins. However, no systematic analysis of their role in the overall degradation of truly short-lived cell proteins has been carried out. Here, the degradation of short-labelled proteins was examined in human fibroblasts by release of trichloroacetic acid-soluble radioactivity. The kinetics of degradation was decomposed into two, corresponding to short- and long-lived proteins, and the effect of proteasomal and lysosomal inhibitors on their degradation, under various growth conditions, was separately investigated. From the degradation kinetics of proteins labelled for various pulse times it can be estimated that about 30% of newly synthesised proteins are degraded with a half-life of approximately 1h. These rapidly degraded proteins should mostly include defective ribosomal products. Deprivation of serum and confluent conditions increased the degradation of the pool of long-lived proteins in fibroblasts without affecting, or affecting to a lesser extent, the degradation of the pool of short-lived proteins. Inhibitors of proteasomes and of lysosomes prevented more than 80% of the degradation of short-lived proteins. It is concluded that, although proteasomes are responsible of about 40-60% of the degradation of short-lived proteins in normal human fibroblasts, lysosomes have also an important participation in the degradation of these proteins. Moreover, in confluent fibroblasts under serum deprivation, lysosomal pathways become even more important than proteasomes in the degradation of short-lived proteins.  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