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.
We have examined interactions between human methylmalonyl CoA mutase and two critical ligands, its cofactor adenosylcobalamin (AdoCbl) and its substrate methylmalonyl CoA, by performing in vitro experiments with preparations of mutase apoenzyme and holoenzyme from normal cultured human fibroblasts. When extracts are prepared from cells grown in medium containing high concentrations of hydroxocobalamin, a precursor of AdoCbl, mutase activity measured in Tris-containing buffers in the absence of added AdoCbl accounts maximally for only 50% of that activity measured in the presence of excess AdoCbl. A similar result is observed when mutase holoenzyme is formed in vitro by incubating cell extracts containing apoenzyme with AdoCbl and removing excess AdoCbl by gel filtration. When such holoenzyme preparations are heated at 45 °C and then assayed for activity, their thermostability is less than that of mutase holoenzyme heated in the presence of excess cofactor, but far greater than that of mutase apoenzyme. Methylmalonyl CoA modulates these enzyme-coenzyme interactions, since mutase holoenzyme formed in Triscontaining buffers is resolved to apoenzyme upon exposure to substrate. Qualitatively different data are obtained when buffers containing cations other than Tris are used. Under these conditions, mutase activity measured in the absence of added AdoCbl accounts for nearly 100% of the activity measured in the presence of excess cofactor, whether holoenzyme is formed in intact cells in culture or in cell extracts in vitro. Furthermore, holoenzyme formed in vitro in potassium phosphate buffer is not resolved to apoenzyme upon exposure to substrate. We suggest that the “holoenzyme” form of mutase obtained and assayed in Tris-containing buffers is that molecular species with only one of its two potential AdoCbl binding sites occupied in a catalytically active fashion, and that other ions can influence markedly the interactions between mutase, AdoCbl, and methylmalonyl CoA. These data are consistent, therefore, with the hypothesis that the dimeric mutase apoenzyme is characterized, under certain conditions, by nonequivalent active sites.  相似文献   

2.
Phosphoglycerate mutases (PGAMs) participate in both the glycolytic and the gluconeogenic pathways in reversible isomerization of 3-phosphoglycerate and 2-phosphoglycerate. PGAMs are members of two distinct protein families: enzymes that are dependent on or independent of the 2,3-bisphosphoglycerate cofactor. We determined the X-ray structure of the monomeric Trypanosoma brucei independent PGAM (TbiPGAM) in its apoenzyme form, and confirmed this observation by small angle X-ray scattering data. Comparing the TbiPGAM structure with the Leishmania mexicana independent PGAM structure, previously reported with a phosphoglycerate molecule bound to the active site, revealed the domain movement resulting from active site occupation. The structure reported here shows the interaction between Asp319 and the metal bound to the active site, and its contribution to the domain movement. Substitution of the metal-binding residue Asp319 by Ala resulted in complete loss of independent PGAM activity, and showed for the first time its involvement in the enzyme's function. As TbiPGAM is an attractive molecular target for drug development, the apoenzyme conformation described here provides opportunities for its use in structure-based drug design approaches. Database Structural data for the Trypanosoma brucei 2,3-bisphosphoglycerate-independent phosphoglycerate mutase (iPGAM) has been deposited with the Research Collaboratory for Structural Bioinformatics (RCSB) Protein Data Bank under code 3NVL.  相似文献   

3.
Frontal and zonal analysis of the chromatography of aspartate aminotransferase (EC2.61.1), pig heart cytosolic enzyme, on Bio-Gel P150 shows that holo- and apoenzyme can dissociate at pH 8.3. Ultracentrifugation and fluorescence depolarization confirm this result. Kinetic analysis of the fluorescence depolarization experiments favors a biphasic phenomenon: a few minutes for the faster one and several hours for the slower one. The apparent dissociation constant is 0.8 muM for the apoenzyme and 0.18 muM for the pyridoxal 5'-phosphate form of the holoenzyme. In the presence of sucrose or 0.1 M L-aspartate or a mixture of 70 mM L-glutamate and 2 mM alpha-ketoglutarate, the holoenzyme is dimeric at concentrations higher than 5 nM. The addition of a mixture of the substrates L-glutamate and alpha-ketoglutarate to a monomeric holoenzyme leads to dimerization. The stability of the dimeric form is in the order: holoenzyme + substrates greater than apoenzyme.  相似文献   

4.
The structures of Leishmania mexicana cofactor-independent phosphoglycerate mutase (Lm iPGAM) crystallised with the substrate 3-phosphoglycerate at high and low cobalt concentrations have been solved at 2.00- and 1.90-Å resolutions. Both structures are very similar and the active site contains both 3-phosphoglycerate and 2-phosphoglycerate at equal occupancies (50%). Lm iPGAM co-crystallised with the product 2-phosphoglycerate yields the same structure. Two Co2+ are coordinated within the active site with different geometries and affinities. The cobalt at the M1 site has a distorted octahedral geometry and is present at 100% occupancy. The M2-site Co2+ binds with distorted tetrahedral geometry, with only partial occupancy, and coordinates with Ser75, the residue involved in phosphotransfer. When the M2 site is occupied, the side chain of Ser75 adopts a position that is unfavourable for catalysis, indicating that this site may not be occupied under physiological conditions and that catalysis may occur via a one-metal mechanism. The geometry of the M2 site suggests that it is possible for Ser75 to be activated for phosphotransfer by H-bonding to nearby residues rather than by metal coordination. The 16 active-site residues of Lm iPGAM are conserved in the Mn-dependent iPGAM from Bacillus stearothermophilus (33% overall sequence identity). However, Lm iPGAM has an inserted tyrosine (Tyr210) that causes the M2 site to diminish in size, consistent with its reduced metal affinity. Tyr210 is present in trypanosomatid and plant iPGAMs, but not in the enzymes from other organisms, indicating that there are two subclasses of iPGAMs.  相似文献   

5.
Production of ATP by the glycolytic pathway in the mammalian pathogenic stage of protists from the genus Trypanosoma is required for the survival of the parasites. Cofactor-independent phosphoglycerate mutase (iPGAM) is particularly attractive as a drug target because it shows no similarity to the corresponding enzyme in humans, and has also been genetically validated as a target by RNAi experiments. It has previously been shown that trypanosomatid iPGAMs require Co(2+) to reach maximal activity, but the biologically relevant metal has remained unclear. In this paper the metal content in the cytosol of procyclic and bloodstream-form T. brucei (analysed by inductively coupled plasma-optical emission spectroscopy) shows that Mg(2+), Zn(2+) and Fe(2+) were the most abundant, whereas Co(2+) was below the limit of detection (<0.035 μM). The low concentration indicates that Co(2+) is unlikely to be the biologically relevant metal, but that instead, Mg(2+) and/or Zn(2+) may assume this role. Results from metal analysis of purified Leishmania mexicana iPGAM by inductively coupled plasma-mass spectrometry also show high concentrations of Mg(2+) and Zn(2+), and are consistent with this proposal. Our data suggest that in vivo cellular conditions lacking Co(2+) are unable to support the maximal activity of iPGAM, but instead maintain its activity at a relatively low level by using Mg(2+) and/or Zn(2+). The physiological significance of these observations is being pursued by structural, biochemical and biophysical studies.  相似文献   

6.
The biosynthesis and assembly of methanol dehydrogenase in bacterium W3A1   总被引:5,自引:0,他引:5  
Bacterium W3A1, a restricted facultative methylotroph, produces a periplasmic methanol dehydrogenase composed of two identical subunits of Mr = 57,300, and two noncovalently bound methoxatin prosthetic groups. A precursor form of Mr = 1,500 larger than the mature subunit was identified among the products of an in vitro translation of total RNA isolated from bacterium W3A1. The precursor form of the protein could not be detected in cells during in vivo pulse-labeling studies, suggesting that the processing of this precursor occurs entirely co-translationally. Whereas the holoenzyme was detectable only as a dimer, removal of the prosthetic group yielded an apoenzyme that could be detected as either a dimeric or monomeric species. After readdition of the purified prosthetic group to the apoenzyme, only the dimeric form of the protein, bearing the cofactor and exhibiting an absorption spectrum similar to that of the holoenzyme, was detected. Neither the mature apoprotein nor the holoenzyme demonstrated any affinity for phospholipid membranes, as assayed by their inability to bind to liposomes. Taken together, these data suggest a scheme of co-translational processing and export of the apoprotein subunits, followed by assembly of the subunits and prosthetic groups in the periplasmic space to form the mature holoenzyme. The suitability of bacterium W3A1, and other methylotrophic bacteria, for use in studies of protein biosynthesis and export, is also discussed.  相似文献   

7.
The apoenzymes of lipoamide dehydrogenase from pig heart and from Pseudomonas fluorescens were prepared at pH 2.7 and pH 4.0, respectively, using a hydrophobic interaction chromatography procedure recently developed for lipoamide dehydrogenase from Azotobacter vinelandii and other flavoproteins [Van Berkel et al. (1988) Eur. J. Biochem. 178, 197-207]. The apoenzyme from pig heart, having 5% of residual activity, shows an equilibrium between the monomeric and dimeric species. Both the yield and the degree of reconstitution of dimeric holoenzyme is 75% of starting material under optimal conditions. The kinetics of reconstitution of pig heart apoenzyme differ slightly from that obtained with the apoenzyme prepared by acid ammonium sulfate precipitation at pH 1.5 [Kalse, J. F. and Veeger, C. (1968) Biochim. Biophys. Acta 159, 244-256]. The apoenzyme from P. fluorescens is in the monomeric state and shows negligible residual activity. The yield and degree of reconstitution of the dimeric holoenzyme is more than 90% of starting material. Reconstitution of the apoenzymes from A. vinelandii and P. fluorescens involves minimally a two-step sequential process. Initial flavin-binding results in regaining of full dichloroindophenol activity, quenching of tryptophan fluorescence and strong increase of FAD fluorescence polarization. In the second step, dimerization occurs as reflected by regain of lipoamide activity, strongly increased FAD fluorescence and increased hyperchroism of the visible absorption spectrum. The kinetics of FAD-induced dimerization are strongly dependent on the apoenzyme used. At 0 degrees C, the monomeric apoenzyme-FAD complex is either stabilized (P. fluorescens) or only transiently detectable (A. vinelandii). Dimerization of P. fluorescens enzyme is strongly stimulated in the presence of NADH.  相似文献   

8.
S Liu  M J Gresser  A S Tracey 《Biochemistry》1992,31(10):2677-2685
The formation of complexes of vanadate with 2-phosphoglycerate and 3-phosphoglycerate have been studied using 51V nuclear magnetic resonance spectroscopy. Signals attributed to two 2,3-diphosphoglycerate analogues, 2-vanadio-3-phosphoglycerate and 2-phospho-3-vanadioglycerate, were detected but were not fully resolved from signals of inorganic vanadate and the anhydride formed between vanadate and the phosphate ester moieties of the individual phosphoglycerates. Equilibrium constants for formation of the two 2,3-bisphosphate analogues were estimated as 2.5 M-1 for 2-vanadio-3-phosphoglycerate and 0.2 M-1 for 2-phospho-3-vanadioglycerate. The results of the binding study are fully consistent with non-cooperativity in the binding of vanadiophosphoglycerate to the two active sites of phosphoglycerate mutase (PGM). 2-Vanadio-3-phosphoglycerate was found to bind to the dephospho form of phosphoglycerate mutase with a dissociation constant of about 1 x 10(-11) M at pH 7 and 7 x 10(-11) M at pH 8. Three signals attributed to histidine residues were observed in the 1H NMR spectrum of phosphoglycerate mutase. Two of these signals and also an additional signal, tentatively attributed to a tryptophan, underwent a chemical shift change when the vanadiophosphoglycerate complex was bound to the enzyme. The results obtained here are in accord with these vanadate-phosphoglycerate complexes being much more potent inhibitors of phosphoglycerate mutase than either monomeric or dimeric vanadate. The dissociation constant of 10(-11) M for 2-vanadio-3-phosphoglycerate is about 4 orders of magnitude smaller than the Km for PGM, a result in accordance with the vanadiophosphoglycerates being transition state analogues for the phosphorylation of PGM by 2,3-diphosphoglycerate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Phosphoglycerate mutases (PGAMs) catalyse the reversible isomerisation of 3-phosphoglycerate and 2-phosphoglycerate, a step of glycolysis. PGAMs can be sub-divided into 2,3-bisphosphoglycerate-dependent (dPGAM) and -independent (iPGAM) enzymes. In plants, phosphoglycerate isomerisation is carried out by cytosolic iPGAM. Despite its crucial role in catabolism, little is known about post-translational modifications of plant iPGAM. In Arabidopsis thaliana, phosphoproteomics analyses have previously identified an iPGAM phosphopeptide where serine 82 is phosphorylated. Here, we show that this phosphopeptide is less abundant in dark-adapted compared to illuminated Arabidopsis leaves. In silico comparison of iPGAM protein sequences and 3D structural modelling of AtiPGAM2 based on non-plant iPGAM enzymes suggest a role for phosphorylated serine in the catalytic reaction mechanism. This is confirmed by the activity (or the lack thereof) of mutated recombinant Arabidopsis iPGAM2 forms, affected in different steps of the reaction mechanism. We thus propose that the occurrence of the S82-phosphopeptide reflects iPGAM2 steady-state catalysis. Based on this assumption, the metabolic consequences of a higher iPGAM activity in illuminated versus darkened leaves are discussed.  相似文献   

10.
A new method is described for the large-scale reversible dissociation of flavoproteins into apoprotein and prosthetic group using hydrophobic-interaction chromatography. Lipoamide dehydrogenase from Azotobacter vinelandii and butyryl-CoA dehydrogenase from Megasphaera elsdenii are selected to demonstrate the usefulness of the method. In contrast to conventional methods, homogeneous preparations of apoproteins in high yields are obtained. The apoproteins show high reconstitutability. The holoenzymes are bound to phenyl-Sepharose CL-4B at neutral pH in the presence of ammonium sulfate. FAD is subsequently removed at pH 3.5-4.0 by addition of high concentrations of KBr. Large amounts of apoenzymes (200-500 mg), showing negligible residual activity, are eluted at neutral pH in the presence of 50% ethylene glycol. The holoenzyme of lipoamide dehydrogenase can be reconstituted while the apoprotein is still bound to the column or the apoenzyme can be isolated in the free state. In both cases the yield and degree of reconstitution of holoenzyme is more than 90% of starting material. Apo-lipoamide-dehydrogenase exists mainly as a monomer in solution and reassociates to the native dimeric structure in the presence of FAD. The apoenzyme is stable for a long period of time when kept in 50% ethylene glycol at -18 degrees C. Steady-state fluorescence-polarization measurements of protein-bound FAD indicate that reconstituted lipoamide dehydrogenase possesses a high stability which is governed by the low dissociation rate constant of the apoenzyme-FAD complex. The holoenzyme of butyryl-CoA dehydrogenase cannot be reconstituted when the apoenzyme is bound to the column. However, stable apoprotein can be isolated in the free state yielding 50-80% of starting material, depending on the immobilization conditions. The coenzyme A ligand present in native holoenzyme is removed during apoprotein preparation. The apoenzyme is relatively stable when kept in 50% ethylene glycol at -18 degrees C. From kinetic and gel filtration experiments it is concluded that the reconstitution reaction of butyryl-CoA dehydrogenase is governed by both the pH-dependent hydrodynamic properties of apoenzyme and the pH-dependent stability of reconstituted enzyme. At pH 7, the apoenzyme is in equilibrium between dimeric and tetrameric forms and reassociates to a native-like tetrameric structure in the presence of FAD. The stability of reconstituted enzyme is strongly influenced by the presence of CoA ligands as shown by fluorescence-polarization measurements. The degree of reconstitution of butyryl-CoA dehydrogenase is more than 80% of the original specific activity under certain conditions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

11.
The conformational stability of holo-lipoamide and apo-lipoamide dehydrogenase from Azotobacter vinelandii was studied by thermoinactivation, unfolding and limited proteolysis. The oxidized holoenzyme is thermostable, showing a melting temperature, tm = 80 degrees C. The thermal stability of the holoenzyme drastically decreases upon reduction. Unlike the oxidized and lipoamide two-electron reduced enzyme species, the NADH four-electron reduced enzyme is highly sensitive to unfolding by urea. Loss of energy transfer from Trp199 to flavin reflects the unfolding of the oxidized holoenzyme by guanidine hydrochloride. Unfolding of the monomeric apoenzyme is a rapid fully reversible process, following a simple two-state mechanism. The oxidized and two-electron reduced holoenzyme are resistant to limited proteolysis by trypsin and endoproteinase Glu-C. Upon cleavage of the apoenzyme or four-electron reduced holoenzyme by both proteases, large peptide fragments (molecular mass greater than 40 kDa) are transiently produced. Sequence studies show that limited trypsinolysis of the NADH-reduced enzyme starts mainly at the C-terminus of Arg391. In the apoenzyme, limited proteolysis by endoproteinase Glu-C starts from the C-terminus at the carboxyl ends of Glu459 and/or Glu435. From crystallographic data it is deduced that the susceptible amino acid peptide bonds are situated near the subunit interface. Thus, these bonds are inaccessible to the proteases in the dimeric enzyme and become accessible after monomerization. It is concluded that reduction of lipoamide dehydrogenase to the four-electron reduced state(s) is accompanied by conformational changes promoting subunit dissociation.  相似文献   

12.
R Hermann  R Jaenicke  N C Price 《Biochemistry》1985,24(8):1817-1821
The reconstitution of the tetrameric phosphoglycerate mutase from bakers' yeast after denaturation in guanidine hydrochloride has been studied. When assays are performed in the presence of trypsin, it is found that reactivation parallels the regain of tetrameric structure. However, in the absence of trypsin, the regain of activity is more rapid, suggesting that monomeric and dimeric intermediates possess partial activity (35% of the value of native enzyme) which is sensitive to trypsin. When reconstitution is studied in the presence of substrates, it is again found that monomeric and dimeric intermediates possess 35% activity. Under these latter conditions, the activity of the monomer but not of the dimer is sensitive to trypsin.  相似文献   

13.
Purified phosphoglycerate mutase from pig skeletal muscle and 2,3-bisphosphoglycerate synthase-phosphatase from pig erythrocytes were hybridized “in vitro”. The hybrid showed a behaviour on electrophoresis and on ion-exchange chromatography similar to that of a naturally occurring enzyme with phosphoglycerate mutase, 2,3-bisphosphoglycerate synthase and 2,3-bisphosphoglycerate phosphatase activities present in pig skeletal and heart muscle. Both the hybrid and the muscle enzyme possess similar activities ratio. From these and previous data it is suggested that the six enzymatic forms with phosphoglycerate mutase, 2,3-bisphosphoglycerate synthase and 2,3-bisphosphoglycerate phosphatase activities detected in mammalian tissues (Carreras et al. 1981, Comp. Biochem. Physiol. 70B, 477–485) result from combination of three subunits (types M, B and E).  相似文献   

14.
The glycolytic enzyme phosphoglycerate mutase (PGM) is of utmost importance for overall cellular metabolism and has emerged as a novel therapeutic target in cancer cells. This enzyme is also conserved in the rapidly proliferating malarial parasite Plasmodium falciparum, which have a similar metabolic framework as cancer cells and rely on glycolysis as the sole energy-yielding process during intraerythrocytic development. There is no redundancy among the annotated PGM enzymes in Plasmodium, and PfPGM1 is absolutely required for the parasite survival as evidenced by conditional knockdown in our study. A detailed comparison of PfPGM1 with its counterparts followed by in-depth structure-function analysis revealed unique attributes of this parasitic protein. Here, we report for the first time the importance of oligomerization for the optimal functioning of the enzyme in vivo, as earlier studies in eukaryotes only focused on the effects in vitro. We show that single point mutation of the amino acid residue W68 led to complete loss of tetramerization and diminished catalytic activity in vitro. Additionally, ectopic expression of the WT PfPGM1 protein enhanced parasite growth, whereas the monomeric form of PfPGM1 failed to provide growth advantage. Furthermore, mutation of the evolutionarily conserved residue K100 led to a drastic reduction in enzymatic activity. The indispensable nature of this parasite enzyme highlights the potential of PfPGM1 as a therapeutic target against malaria, and targeting the interfacial residues critical for oligomerization can serve as a focal point for promising drug development strategies that may not be restricted to malaria only.  相似文献   

15.
In pig skeletal muscle exist four enzymes with 2,3-bisphosphoglycerate phosphatase activity. Two of them (forms I-A and I-C) are multi-functional enzymes which, in addition to the phosphatase activity, possess 2,3-bisphosphoglycerate synthase and phosphoglycerate mutase activities. The other two enzyme forms (II-A and II-B) only show the phosphatase activity. The four enzymes differ in substrate specificity. Form I-C is highly specific for glycerate 2,3-P2; form I-A also hydrolyzes the monophosphoglycerates and forms II-A and II-B are specific for phosphoester bonds adjacent to a C-1 carboxylic group. The enzymes possess similar Km, Kcat and optimum pH value, but they are differently inhibited by the reaction products. They are also differently affected by glycolate-2-P (their main activator) and by other modifiers. Probably form I-A, which corresponds to M-type phosphoglycerate mutase, is the main enzyme implicated in the breakdown of glycerate 2,3-P2 in pig muscle.  相似文献   

16.
Cofactor-independent phosphoglycerate mutase (iPGAM) is essential for the growth of C. elegans but is absent from humans, suggesting its potential as a drug target in parasitic nematodes such as Brugia malayi, a cause of lymphatic filariasis (LF). iPGAM''s active site is small and hydrophilic, implying that it may not be druggable, but another binding site might permit allosteric inhibition. As a comprehensive assessment of iPGAM''s druggability, high-throughput screening (HTS) was conducted at two different locations: ∼220,000 compounds were tested against the C. elegans iPGAM by Genzyme Corporation, and ∼160,000 compounds were screened against the B. malayi iPGAM at the National Center for Drug Screening in Shanghai. iPGAM''s catalytic activity was coupled to downstream glycolytic enzymes, resulting in NADH consumption, as monitored by a decline in visible-light absorbance at 340 nm. This assay performed well in both screens (Z′-factor >0.50) and identified two novel inhibitors that may be useful as chemical probes. However, these compounds have very modest potency against the B. malayi iPGAM (IC50 >10 µM) and represent isolated singleton hits rather than members of a common scaffold. Thus, despite the other appealing properties of the nematode iPGAMs, their low druggability makes them challenging to pursue as drug targets. This study illustrates a “druggability paradox” of target-based drug discovery: proteins are generally unsuitable for resource-intensive HTS unless they are considered druggable, yet druggability is often difficult to predict in the absence of HTS data.  相似文献   

17.
Effects of temperature and monovalent cations on the activity and the quaternary structure of tryptophanase of Escherichia coli were studied. The conversion of the apoenzyme into the active holoenzyme was attained at 30 degrees C in Tris-HCl buffer (pH 8.0) containing pyridoxal-P and K+, while no conversion occurred at 5 degrees C. The active holoenzyme thus formed was stable even at 5 degrees C, as long as the cation was present. When K+ was absent, however, the active enzyme gradually lost the activity upon chilling to 5 degrees C. The HPLC gel filtration analysis of the active holoenzyme and the low temperature-inactivated enzyme species revealed that the tetrameric holoenzyme dissociated into the dimeric apoenzyme concomitant with the low temperature-induced inactivation at 5 degrees C. The results of HPLC experiments together with other available evidence also suggest that the inactive tetrameric holoenzyme was first formed from the dimeric apoenzyme and pyridoxal-P prior to the formation of the active holoenzyme and that the cation promoted the conversion of the inactive holoenzyme into the active holoenzyme rather than being involved in the conversion of the apoenzyme and pyridoxal-P into the holoenzyme. Among various cations tested for the above effects, NH4+ exhibited the largest effect and K+ the second.  相似文献   

18.
19.
  • Glycolysis is a central metabolic pathway that provides energy and products of primary metabolites. 2,3‐Biphosphoglycerate‐independent phosphoglycerate mutase (iPGAM) is a key enzyme that catalyses the reversible interconversion of 3‐phosphoglycerate (3‐PGA) to 2‐phosphoglycerate (2‐PGA) in glycolysis. Low temperature is a common abiotic stress in rice production. However, the mechanism for rice iPGAM genes is not fully understood at low temperature.
  • In this study, the rice mutant tcm12, with chlorosis, malformed chloroplasts and impaired photosynthesis, was grown at a low temperature (<20 °C) to the three‐leaf stage, while the normal phenotype at 32 °C was used. Chlorophyll fluorescence analysis and transmission electron microscopy were used to examine features of the tcm12 mutant. The inheritance behaviour and function of TCM12 were then analysed thorough map‐based cloning, transgenic complementation and subcellular localisation.
  • The thermo‐sensitive chlorosis phenotype was caused by a single nucleotide mutation (T→C) on the fifth exon of TCM12 (LOC_Os12g35040) encoding iPGAM, localised to both nucleus and membranes. In addition, TCM12 was constitutively expressed, and its disruption resulted in down‐regulation of some genes associated with chlorophyll biosynthesis and photosynthesis at low temperatures (20 °C).
  • This is the first report of the involvement of rice iPGAM gene in chlorophyll synthesis, photosynthesis and chloroplast development, providing new insights into the mechanisms underlying early growth of rice at low temperatures.
  相似文献   

20.
The anthranilate phosphoribosyltransferase from Sulfolobus solfataricus (ssAnPRT) forms a homodimer with a hydrophobic subunit interface. To elucidate the role of oligomerisation for catalytic activity and thermal stability of the enzyme, we loosened the dimer by replacing two apolar interface residues with negatively charged residues (mutations I36E and M47D). The purified double mutant I36E+M47D formed a monomer with wild-type catalytic activity but reduced thermal stability. The single mutants I36E and M47D were present in a monomer-dimer equilibrium with dissociation constants of about 1 μM and 20 μM, respectively, which were calculated from the concentration-dependence of their heat inactivation kinetics. The monomeric form of M47D, which is populated at low subunit concentrations, was as thermolabile as monomeric I36E+M47D. Likewise, the dimeric form of I36E, which was populated at high subunit concentrations, was as thermostable as dimeric wild-type ssAnPRT. These findings show that the increased stability of wild-type ssAnPRT compared to the I36E+M47D double mutant is not caused by the amino acid exchanges per se but by the higher intrinsic stability of the dimer compared to the monomer. In accordance with the negligible effect of the mutations on catalytic activity and stability, the X-ray structure of M47D contains only minor local perturbations at the dimer interface. We conclude that the monomeric double mutant resembles the individual wild-type subunits, and that ssAnPRT is a dimer for stability but not for activity reasons.  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