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.
2.
Our previous research characterized two phosphoenolpyruvate (PEP) carboxylase (PEPC) isoforms (PEPC1 and PEPC2) from developing castor oil seeds (COS). The association of a shared 107-kD subunit (p107) with an immunologically unrelated bacterial PEPC-type 64-kD polypeptide (p64) leads to marked physical and kinetic differences between the PEPC1 p107 homotetramer and PEPC2 p107/p64 heterooctamer. Here, we describe the production of antiphosphorylation site-specific antibodies to the conserved p107 N-terminal serine-6 phosphorylation site. Immunoblotting established that the serine-6 of p107 is phosphorylated in COS PEPC1 and PEPC2. This phosphorylation was reversed in vitro following incubation of clarified COS extracts or purified PEPC1 or PEPC2 with mammalian protein phosphatase type 2A and is not involved in a potential PEPC1 and PEPC2 interconversion. Similar to other plant PEPCs examined to date, p107 phosphorylation increased PEPC1 activity at pH 7.3 by decreasing its K(m)(PEP) and sensitivity to L-malate inhibition, while enhancing glucose-6-P activation. By contrast, p107 phosphorylation increased PEPC2's K(m)(PEP) and sensitivity to malate, glutamic acid, and aspartic acid inhibition. Phosphorylation of p107 was promoted during COS development (coincident with a >5-fold increase in the I(50) [malate] value for total PEPC activity in desalted extracts) but disappeared during COS desiccation. The p107 of stage VII COS became fully dephosphorylated in planta 48 h following excision of COS pods or following 72 h of dark treatment of intact plants. The in vivo phosphorylation status of p107 appears to be modulated by photosynthate recently translocated from source leaves into developing COS.  相似文献   

3.
Phosphoenolpyruvate carboxylase (PEPC) is believed to play an important role in producing malate as a substrate for fatty acid synthesis by leucoplasts of the developing castor oilseed (COS) endosperm. Two kinetically distinct isoforms of COS PEPC were resolved by gel filtration chromatography and purified. PEPC1 is a typical 410-kDa homotetramer composed of 107-kDa subunits (p107). In contrast, PEPC2 exists as an unusual 681-kDa hetero-octamer composed of the same p107 found in PEPC1 and an associated 64-kDa polypeptide (p64) that is structurally and immunologically unrelated to p107. Relative to PEPC1, PEPC2 demonstrated significantly enhanced thermal stability and a much lower sensitivity to allosteric activators (Glc-6-P, Glc-1-P, Fru-6-P, glycerol-3-P) and inhibitors (Asp, Glu, malate) and pH changes within the physiological range. Nondenaturing PAGE of clarified extracts followed by in-gel PEPC activity staining indicated that the ratio of PEPC1:PEPC2 increases during COS development such that only PEPC1 is detected in mature COS. Dissimilar developmental profiles and kinetic properties support the hypotheses that (i) PEPC1 functions to replenish dicarboxylic acids consumed through transamination reactions required for storage protein synthesis, whereas (ii) PEPC2 facilitates PEP flux to malate in support of fatty acid synthesis. Interestingly, the respective physical and kinetic properties of COS PEPC1 and PEPC2 are remarkably comparable with those of the homotetrameric low M(r) Class 1 and heteromeric high M(r) Class 2 PEPC isoforms of unicellular green algae.  相似文献   

4.
The phosphoenolpyruvate carboxylase (PEPC) interactome of developing castor oil seed (COS; Ricinus communis) endosperm was assessed using coimmunopurification (co-IP) followed by proteomic analysis. Earlier studies suggested that immunologically unrelated 107-kD plant-type PEPCs (p107/PTPC) and 118-kD bacterial-type PEPCs (p118/BTPC) are subunits of an unusual 910-kD hetero-octameric class 2 PEPC complex of developing COS. The current results confirm that a tight physical interaction occurs between p118 and p107 because p118 quantitatively coimmunopurified with p107 following elution of COS extracts through an anti-p107-IgG immunoaffinity column. No PEPC activity or immunoreactive PEPC polypeptides were detected in the corresponding flow-through fractions. Although BTPCs lack the N-terminal phosphorylation motif characteristic of PTPCs, Pro-Q Diamond phosphoprotein staining, immunoblotting with phospho-serine (Ser)/threonine Akt substrate IgG, and phosphate-affinity PAGE established that coimmunopurified p118 was multiphosphorylated at unique Ser and/or threonine residues. Tandem mass spectrometric analysis of an endoproteinase Lys-C p118 peptide digest demonstrated that Ser-425 is subject to in vivo proline-directed phosphorylation. The co-IP of p118 with p107 did not appear to be influenced by their phosphorylation status. Because p118 phosphorylation was unchanged 48 h following elimination of photosynthate supply due to COS depodding, the signaling mechanisms responsible for photosynthate-dependent p107 phosphorylation differ from those controlling p118's in vivo phosphorylation. A 110-kD PTPC coimmunopurified with p118 and p107 when depodded COS was used. The plastidial pyruvate dehydrogenase complex (PDC(pl)) was identified as a novel PEPC interactor. Thus, a putative metabolon involving PEPC and PDC(pl) could function to channel carbon from phosphoenolpyruvate to acetyl-coenzyme A and/or to recycle CO(2) from PDC(pl) to PEPC.  相似文献   

5.
Murmu J  Plaxton WC 《Planta》2007,226(5):1299-1310
Phosphoenolpyruvate carboxylase (PEPC, EC 4.1.1.31) protein kinase (PPCK) was purified ∼1,500-fold from developing castor oil seeds (COS). Gel filtration and immunoblotting with anti-(rice PPCK2)-immune serum indicated that this Ca2+-insensitive PPCK exists as a 31-kDa monomer. COS PPCK-mediated rephosphorylation of the 107-kDa subunit (p107) of COS PEPC1 (K m = 2.2 μM) activated PEPC1 by ∼80% when assayed under suboptimal conditions (pH 7.3, 0.2 mM PEP, and 0.125 mM malate). COS PPCK displayed remarkable selectivity for phosphorylating COS PEPC1 (relative to tobacco, sorghum, or maize PEPCs), exhibited a broad pH-activity optima of ∼pH 8.5, and at pH 7.3 was activated 40–65% by 1 mM PEP, or 10 mM Gln or Asn, but inhibited 65% by 10 mM L-malate. The possible control of COS PPCK by disulfide-dithiol interconversion was suggested by its rapid inactivation and subsequent reactivation when incubated with oxidized glutathione and then dithiothreitol. In vitro PPCK activity correlated with in vivo p107 phosphorylation status, with both peaking in mid-cotyledon to full-cotyledon developing COS. Notably, PPCK activity and p107 phosphorylation of developing COS were eliminated following pod excision or prolonged darkness of intact plants. Both effects were fully reversed 12 h following reillumination of darkened plants. These results implicate a direct relationship between the up-regulation of COS PPCK and p107 phosphorylation during the recommencement of photosynthate delivery from illuminated leaves to the non-photosynthetic COS. Overall, the results support the hypothesis that PEPC and PPCK participate in the control of photosynthate partitioning into C-skeletons needed as precursors for key biosynthetic pathways of developing COS. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

6.
7.
8.
Two isozymes (AIV I and AIV II) of soluble acid invertase (EC 3.2.1.26) were purified from Japanese pear fruit through procedures including (NH(4))(2)SO(4) precipitating, DEAE-Sephacel column chromatography, Concanavalin A (ConA)-Sepharose affinity chromatography, hydroxyapatite column chromatography and Mono Q HR 5/5 column chromatography. The specific activities of purified AIV I and AIV II were 2670 and 2340 (nkat/mg protein), respectively. AIV I was a monomeric enzyme of 80 kDa, while AIV II may be also a monomeric enzyme, which is easy to be cleaved to 52 kDa and 34 kDa polypeptide during preparation by SDS-PAGE. The Km values for sucrose of AIV I and AIV II were 3.33 and 4.58 mM, respectively, and optimum pH of both enzyme activities was pH 4.5.  相似文献   

9.
1. Washed suspensions of two Achromobacter species (G2 and 2L), capable of growth upon 2- and 3-hydroxypyridine respectively as sources of C and N, rapidly oxidized their growth substrate pyridine-2,5-diol (2,5-dihydroxypyridine) and the putative ring-cleavage product maleamate without a lag. Suspensions derived from fumarate plus (NH(4))(2)SO(4) cultures were unable to do so. 2. Extracts of both bacteria oxidized pyridine-2,5-diol with the stoicheiometry of an oxygenase forming 1mol of NH(3)/mol of substrate. 3. Heat-treated extracts, however, formed maleamate and formate with little free NH(3). 4. The conversion of maleamate into maleate plus NH(3) by extracts of strain 2L, fractionated with (NH(4))(2)SO(4), and the metabolism of maleamate and maleate to fumarate by extracts of both strains demonstrated the existence of the enzymes catalysing each reaction of the maleamate pathway in these bacteria. 5. The pyridine-2,5-diol dioxygenase (mol.wt. approx. 340000) in extracts of these Achromobacter species required Fe(2+) (1.7mum) to restore full activity after dialysis or treatment with chelating agents; the enzyme from strain 2L also had a specific requirement for l-cysteine (6.7mm), which could not be replaced by GSH or dithiothreitol. 6. The oxygenase was strongly inhibited in a competitive manner by the isomeric pyridine-2,3- and -3,4-diols.  相似文献   

10.
Reticulocytes contain a nonlysosomal proteolytic pathway that requires ATP and ubiquitin. By DEAE chromatography and gel filtration, we were able to fractionate the ATP-dependent system into a 30-300-kDa fraction that catalyzes the ATP-dependent conjugation of ubiquitin to substrates ("Conjugation Fraction") and a high mass fraction (greater than 450 kDa) necessary for hydrolysis of the conjugated proteins. The latter contains two distinct proteases. One protease is unusually large, approximately 1500 kDa, and degrades proteins only when ATP and the conjugating fractions are added. This activity precipitates at 0-38% (NH4)2SO4 saturation and is essential for ATP-dependent proteolysis. Like crude extracts, it is labile in the absence of nucleotides and is inhibited by heparin, poly(Glu-Ala-Tyr), 3,4-dichloroisocoumarin, hemin, decavanadate, N-ethylmaleimide, and various peptide chloromethyl ketones. It lacks amino-peptidase and insulin-degrading activities and does not require tRNA for activity. The ubiquitin-conjugate degrading enzyme, which we suggest be named UCDEN, is inactive against substrates that cannot undergo ubiquitin conjugation. The smaller protease (670 kDa), which precipitates at 40-80% (NH4)2SO4 saturation, does not require ATP or ubiquitin and is therefore not required for ATP-dependent proteolysis. It is stimulated by N-ethylmaleimide and 3,4-dichloroisocoumarin and is stable at 37 degrees C. It hydrolyzes fluorometric tetrapeptides and proteins, including proteins which cannot be conjugated to ubiquitin. Thus, reticulocytes contain two large cytosolic proteases: one is essential for the degradation of ubiquitin conjugates, while the function of the other is uncertain.  相似文献   

11.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

12.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

13.
Phosphoenolpyruvate carboxykinase (PEPCK; EC 4.1.1.49) catalyses the reversible decarboxylation of oxaloacetate to phosphoenolpyruvate in the gluconeogenic production of sugars from storage lipids in germinating oil seeds. The enzyme is quite susceptible to limited proteolysis during extraction. Immunoblotting was used to diagnose unwanted in vitro proteolytic activity against PEPCK from germinating castor oil seeds (COS) by following the disappearance of its native 74-kDa subunit and concomitant appearance of a truncated 64-kDa polypeptide. Alkaline pH and the inclusion of thiol protease inhibitors effectively prevented COS PEPCK proteolysis during incubation of clarified COS extracts at 4°C. The carboxylating and decarboxylating activities and concentration of non-proteolyzed COS PEPCK were investigated during germination. This is the first report in which both activities catalyzed by PEPCK were measured in vitro during a whole developmental process. Carboxylating activity and the level of immunoreactive 74-kDa PEPCK polypeptides rapidly increased in parallel to maximal values by day 5 and then significantly declined over the subsequent 2 days. In contrast, decarboxylating PEPCK activity was much higher over the 7 days of growth examined. In addition, the effect on PEPCK activity while changing the endosperm from aerobic (when gluconeogenesis predominates in the tissue) to anaerobic conditions (where the tissue becomes glycolytic) was studied. While PEPCK decarboxylating activity remained almost constant, carboxylating activity declined to undetectable levels in response to anaerobiosis. These and the developmental profile results suggest that COS PEPCK may be subject to a mechanism of post-translation control that selectively inhibits the carboxylating, but not the decarboxylating activity.  相似文献   

14.
PEPC [PEP (phosphoenolpyruvate) carboxylase] is a tightly controlled anaplerotic enzyme situated at a pivotal branch point of plant carbohydrate metabolism. Two distinct oligomeric PEPC classes were discovered in developing COS (castor oil seeds). Class-1 PEPC is a typical homotetramer of 107?kDa PTPC (plant-type PEPC) subunits, whereas the novel 910-kDa Class-2 PEPC hetero-octamer arises from a tight interaction between Class-1 PEPC and 118?kDa BTPC (bacterial-type PEPC) subunits. Mass spectrometric analysis of immunopurified COS BTPC indicated that it is subject to in vivo proline-directed phosphorylation at Ser425. We show that immunoblots probed with phosphorylation site-specific antibodies demonstrated that Ser425 phosphorylation is promoted during COS development, becoming maximal at stage IX (maturation phase) or in response to depodding. Kinetic analyses of a recombinant, chimaeric Class-2 PEPC containing phosphomimetic BTPC mutant subunits (S425D) indicated that Ser425 phosphorylation results in significant BTPC inhibition by: (i) increasing its Km(PEP) 3-fold, (ii) reducing its I50 (L-malate and L-aspartate) values by 4.5- and 2.5-fold respectively, while (iii) decreasing its activity within the physiological pH range. The developmental pattern and kinetic influence of Ser425 BTPC phosphorylation is very distinct from the in vivo phosphorylation/activation of COS Class-1 PEPC's PTPC subunits at Ser11. Collectively, the results establish that BTPC's phospho-Ser425 content depends upon COS developmental and physiological status and that Ser425 phosphorylation attenuates the catalytic activity of BTPC subunits within a Class-2 PEPC complex. To the best of our knowledge, this study provides the first evidence for protein phosphorylation as a mechanism for the in vivo control of vascular plant BTPC activity.  相似文献   

15.
An N-acetylglucosamine-binding lectin with a molecular mass of 32kDa was isolated from fresh sclerotia of the edible mushroom Pleurotus tuber-regium. Its N-terminal sequence exhibited some similarity to that of Agaricus bisporus lectin. The isolation procedure was simple, involving (NH(4))(2)SO(4) precipitation, ion exchange chromatography on DEAE-cellulose, affinity chromatography on N-acetyl-D-glucosamine-agarose, and gel filtration by fast protein liquid chromatography on Superdex 75. The lectin exhibited hemagglutinating activity toward trypsinized rabbit erythrocytes but not toward untrypsinized rabbit erythrocytes.  相似文献   

16.
17.
18.
Interaction of ras oncogene product p21 with guanine nucleotides   总被引:2,自引:0,他引:2  
The nucleotide exchange reaction was observed with purified ras oncogene product p21 overproduced in Escherichia coli (Hattori, S. et al. (1985) Mol. Cell Biol. 5, 1449-1455) under various conditions. (NH4)2SO4 increased the rate of dissociation of bound GDP from c-rasH and v-rasH p21. The dissociation kinetics were those of a first order reaction, and there was a linear relationship between the rate constant and the (NH4)2SO4 concentration. At any concentration of (NH4)2SO4, the exchange rate was faster with v-rasH p21 than that with c-rasH p21. EDTA and (NH4)2SO4 synergetically stimulated the dissociation reaction. Nucleotide-free p21 was prepared by gel filtration on Sephadex G-25 in the presence of 5 mM EDTA and 200 mM (NH4)2SO4 at room temperature. The free p21 was quite thermolabile, but the addition of GDP or GTP completely protected p21 from thermal inactivation. The dissociation constants for GDP and GTP were determined with free p21 to be 8.9 and 8.2 nM, respectively, for v-rasH p21, and 1.0 and 2.6 nM for c-rasH p21. In the presence of 200 mM (NH4)2SO4, these dissociation constants increased 3- to 12-fold.  相似文献   

19.
Soluble peroxidase (POD) from oil palm leaf was purified by (NH(4))(2)SO(4) precipitation, anion exchange chromatography and molecular exclusion chromatography. The purification grade obtained was 429 yielding 54% of the enzyme activity. Electrophoresis of purified enzyme under denatured conditions revealed M(r) of 48+/-2 kDa. It has an optimum pH of 5 and it exhibited very high pH and thermal stabilities. K(m) for guaiacol, ABTS and pyrogallol were 3.96, 1 and 0.84 mM, respectively. Immunocytochemical localization studies showed that soluble POD was mainly located in the vascular bundles and epidermis of leaf.  相似文献   

20.
Two different extracellular proteases, protease I (P-I), an alkaline protease, and protease II (P-II) a neutral protease, from Aspergillus flavus var. columnaris were partially purified by using (NH(4))(2)SO(4) precipitation, diethylaminoethyl-Sephadex A-50 chromatography, carboxymethylcellulose CM-52 chromatography, and Sephadex G-100 gel filtration. The degree of purity was followed using polyacrylamide gel electrophoresis. The activity of P-I was completely inhibited by 0.1 mM phenylmethylsulfonyl fluoride, and that of P-II was completely inhibited by 1 mM ethylenediaminetetraacetate. By using these inhibitors with extracts of wheat bran koji, the proportions of total activity that could be assigned to P-I and P-II were 80 and 20%, respectively. This compared favorably with activities estimated by using polyacrylamide gel electrophoresis slices (82 and 18%, respectively). Extracts from factory-run soybean koji gave comparable results. Both enzymes demonstrated maximum activity at 50 to 55 degrees C and only small changes in activity between pH 6 and 11. For P-I, activity was somewhat higher from pH 8.0 to 11.0, whereas for P-II it was somewhat higher from pH 6 to 9. In the presence of 18% NaCl, the activities of both P-I and P-II dropped by approximately 90 and 85%, respectively. P-I was inferred to possess aminopeptidase activity since it could hydrolyze l-leucyl-p-nitroanilide hydrochloride. P-II was devoid of such activity. The ramifications of the results for factory-produced soy sauce koji are discussed.  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