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.
The plastidic glyceraldehyde‐3‐phosphate dehydrogenase (GAPDH) catalyzes the only reductive step in the Calvin cycle and exists as different forms of which GapC1 enzyme is present in chromalveolates, such as diatoms. Biochemical studies on diatoms are still fragmentary, and, thus, in this report, GAPDH from the freshwater diatom Asterionella formosa Hassall has been purified and kinetically characterized. It is a homotetrameric enzyme with a molecular mass of ~150 ± 15 kDa. The enzyme showed Michaelis–Menten kinetics with respect to both cofactors, NADPH and NADH, with a 16‐fold greater catalytic constant for NADPH. The Km for NADPH was 140 μM, the lowest affinity reported, while the catalytic constant, 815 s?1, is the highest reported. The Km for NADH was 93 μM, and the catalytic constant was 50 s?1, both are similar to reported values for other types of GAPDH. The GapC1 enzyme, like the Chlamydomonas reinhardtii A4 GAPDH, exhibits a cooperative behavior toward the substrate, 1,3‐bisphosphoglyceric acid (BPGA), with both cofactors. Mass spectrometry analysis showed that when GapC1 enzyme was purified without reducing agents, it copurified with a small protein with a mass of 8.2 kDa. This protein was recognized by antibodies against CP12. When associated with this protein, GAPDH displayed a lag that disappeared upon incubation with reducing agent in the presence of either BPGA or NADPH as a consequence of dissociation of the GAPDH/CP12 complex. Thus, as in other species of algae and higher plants, regulation of GapC1 enzyme in A. formosa may occur through association‐dissociation processes linked to dark‐light transitions.  相似文献   

2.
Marri L  Trost P  Pupillo P  Sparla F 《Plant physiology》2005,139(3):1433-1443
Calvin cycle enzymes glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoribulokinase (PRK) form together with the regulatory peptide CP12 a supramolecular complex in Arabidopsis (Arabidopsis thaliana) that could be reconstituted in vitro using purified recombinant proteins. Both enzyme activities were strongly influenced by complex formation, providing an effective means for regulation of the Calvin cycle in vivo. PRK and CP12, but not GapA (A(4) isoform of GAPDH), are redox-sensitive proteins. PRK was reversibly inhibited by oxidation. CP12 has no enzymatic activity, but it changed conformation depending on redox conditions. GapA, a bispecific NAD(P)-dependent dehydrogenase, specifically formed a binary complex with oxidized CP12 when bound to NAD. PRK did not interact with either GapA or CP12 singly, but oxidized PRK could form with GapA/CP12 a stable ternary complex of about 640 kD (GapA/CP12/PRK). Exchanging NADP for NAD, reducing CP12, or reducing PRK were all conditions that prevented formation of the complex. Although GapA activity was little affected by CP12 alone, the NADPH-dependent activity of GapA embedded in the GapA/CP12/PRK complex was 80% inhibited in respect to the free enzyme. The NADH activity was unaffected. Upon binding to GapA/CP12, the activity of oxidized PRK dropped from 25% down to 2% the activity of the free reduced enzyme. The supramolecular complex was dissociated by reduced thioredoxins, NADP, 1,3-bisphosphoglycerate (BPGA), or ATP. The activity of GapA was only partially recovered after complex dissociation by thioredoxins, NADP, or ATP, and full GapA activation required BPGA. NADP, ATP, or BPGA partially activated PRK, but full recovery of PRK activity required thioredoxins. The reversible formation of the GapA/CP12/PRK supramolecular complex provides novel possibilities to finely regulate GapA ("non-regulatory" GAPDH isozyme) and PRK (thioredoxin sensitive) in a coordinated manner.  相似文献   

3.
1. Long‐term data on the meteorology, hydrology, physicochemistry and plankton of a reservoir and its tributaries in SE Germany run from 1976 until now. This dimictic reservoir changed from mesotrophic to eutrophic in the 1970s, remained eutrophic in the 1980s and returned to the mesotrophic state after a sharp reduction in P loading in 1990. 2. Phytoplankton biomass reaches an annual maximum in spring and consists almost entirely of diatoms. While Asterionella formosa was dominant until 1990, Aulacoseira subarctica became more frequent at the end of the 1990s and was particularly abundant in years with short winters. 3. Statistical analyses suggested that these changes were triggered primarily by the mild winters that were frequent after 1988. Climate‐related hydrophysical variables and the initial biomass of the diatoms at the beginning of the year, considered as an ‘inoculum’, were identified as most important. These variables explained 39% of the total variance of the relative abundance, whereas the change in trophic conditions was responsible for about 20%. 4. The absolute and relative abundance of A. subarctica was positively related to short ice cover, early ice‐out and a long‐lasting spring circulation. Owing to its physiological traits, and particularly its ability to survive under low‐light conditions, A. subarctica benefitted from short, mild winters. Under such conditions, it could sustain or establish a high initial biomass, whereas the concentrations of the other diatoms decreased over winter. However, this advantage may be lost if further warming causes an early onset of summer stratification. Because of its low population growth rate and requirement for high turbulence, A. subarctica needs long, cold springs to exploit the improved starting conditions and to become abundant. 5. In contrast to A. subarctica, A. formosa required a substantial soluble reactive phosphorus supply to compete successfully. The eutrophic conditions until 1990 were the prerequisite for its mass growth under low‐light and low‐temperature conditions during the spring. After reduction in P concentration from 1990, A. formosa declined and other diatom species became more abundant. 6. These other diatoms may be viewed as ‘stopgaps’ when conditions were not favourable for A. subarctica or A. formosa. Diatoma elongatum exploited brief circulation periods in years with low P loading. Synedra acus and Fragilaria crotonensis, because of their poor competitive ability at low light intensity, reached high density in the upper water column in the transitional period between spring circulation and summer stratification. 7. Our study suggests that climate‐related variables have crucial impacts on the spring phytoplankton dynamics of deep stratified waterbodies. They can mask the consequences of changes in the trophic conditions and, corresponding to the functional traits of the different phytoplankton species, also decisively control their relative abundances. In this reservoir, the warmer winters and prolonged spring circulations did not only lead to high phytoplankton biomass (despite considerably reduced nutrient loads) but also cause a marked shift in the diatom assemblage during the spring bloom.  相似文献   

4.
The activity of glyceraldehyde-3-phosphate dehydrogenase (GAPDH) embedded in the phosphoribulokinase (PRK).GAPDH.CP12 complex was increased 2-3-fold by reducing agents. This occurred by interaction with PRK as the cysteinyl sulfhydryls (4 SH/subunit) of GAPDH within the complex were unchanged whatever the redox state of the complex. But isolated GAPDH was not activated. Alkylation plus mass spectrometry also showed that PRK had one disulfide bridge and three SH groups per monomer in the active oxidized complex. Reduction disrupted this disulfide bridge to give 2 more SH groups and a much more active enzyme. We assessed the kinetics and dynamics of the interactions between PRK and GAPDH/CP12 using biosensors to measure complex formation in real time. The apparent equilibrium binding constant for GAPDH/CP12 and PRK was 14 +/- 1.6 nm for oxidized PRK and 62 +/- 10 nm for reduced PRK. These interactions were neither pH- nor temperature-dependent. Thus, the dynamics of PRK.GAPDH.CP12 complex formation and GAPDH activity are modulated by the redox state of PRK.  相似文献   

5.
The glyceraldehyde-3-phosphate dehydrogenase (GAPDH) in the chloroplast of Chlamydomonas reinhardtii is part of a complex that also includes phosphoribulokinase (PRK) and CP12. We identified two residues of GAPDH involved in protein-protein interactions in this complex, by changing residues K128 and R197 into A or E. K128A/E mutants had a Km for NADH that was twice that of the wild type and a lower catalytic constant, whatever the cofactor. The kinetics of the mutant R197A were similar to those of the wild type, while the R197E mutant had a lower catalytic constant with NADPH. Only small structural changes near the mutation may have caused these differences, since circular dichroism and fluorescence spectra were similar to those of wild-type GAPDH. Molecular modelling of the mutants led to the same conclusion. All mutants, except R197E, reconstituted the GAPDH-CP12 subcomplex. Although the dissociation constants measured by surface plasmon resonance were 10-70-fold higher with the mutants than with wild-type GAPDH and CP12, they remained low. For the R197E mutation, we calculated a 4 kcal/mol destabilizing effect, which may correspond to the loss of the stabilizing effect of a salt bridge for the interaction between GAPDH and CP12. All the mutant GAPDH-CP12 subcomplexes failed to interact with PRK and to form the native complex. The absence of kinetic changes of all the mutant GAPDH-CP12 subcomplexes, compared to wild-type GAPDH-CP12, suggests that mutants do not undergo the conformation change essential for PRK binding.  相似文献   

6.
7.
Two Calvin Cycle enzymes, NAD(P)-dependent glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoribulokinase (PRK) form a multiprotein complex with CP12, a small intrinsically-unstructured protein. Under oxidizing conditions, association with CP12 confers redox-sensitivity to the otherwise redox-insensitive A isoform of GAPDH (GapA) and provides an additional level of down-regulation to the redox-regulated PRK. To determine if CP12-mediated regulation is specific for GAPDH and PRK in vivo, a high molecular weight complex containing CP12 was isolated from tobacco chloroplasts and leaves and its protein composition was characterized. Gel electrophoresis and immunoblot analyses after separation of stromal proteins by size fractionation verified that the GAPDH (both isoforms) and PRK co-migrated with CP12 in dark- but not light-adapted chloroplasts. Nano-liquid-chromatography-mass-spectrometry of the isolated complex identified only CP12, GAPDH and PRK. Since nearly all of the CP12 from darkened chloroplasts migrates with GADPH and PRK as a high molecular mass species, these data indicate that the tight association of tobacco CP12 with GAPDH and PRK is specific and involves no other Calvin Cycle enzymes.  相似文献   

8.
1. Historical nutrient changes in Grasmere were investigated using a 300‐year record derived from six sediment cores. One core was investigated at high resolution for diatoms, total sedimentary phosphorus, and loss‐on‐ignition (LOI), and was dated using 210Pb and 137Cs. Six other cores were scanned for magnetic susceptibility, diatoms and LOI to confirm the stratigraphic integrity of the primary record. 2. A rise in nutrient levels occurred after 1855 AD. This event was marked by a shift away from benthic diatom assemblages and a rise in Asterionella formosa. The onset of eutrophication from 1855 corresponds to the expansion of the local and tourist population in the area. 3. The replacement of A. formosa with Cyclotella spp. ca 1945–65 indicates reduced nutrient loads, possibly because of enhanced flushing brought about by the seasonal rainfall distribution. 4. After 1965 a step‐wise increase in both absolute and relative amounts of Asterionella was found. High sedimentary P and diatom inferred TP confirmed the high nutrient loading of the lake. Nutrient increase is attributable to problems with the Grasmere village sewage system and the installation of a wastewater treatment works (WwTW) on the River Rothay in 1971. Modifications to the WwTW in 1982 caused an initial improvement, but have not led to a full recovery to pre‐1965 ecological conditions. 5. The diatom record indicates a further improvement after 1990 by a return toward Achnanthes minutissima. 6. The sedimentary archive of sensitive sites provides important benchmarks against which to judge the attainment of water quality targets.  相似文献   

9.
10.
11.
CP12, a small intrinsically unstructured protein, plays an important role in the regulation of the Calvin cycle by forming a complex with phosphoribulokinase (PRK) and glyceraldehyde-3-phosphate dehydrogenase (GAPDH). An extensive search in databases revealed 129 protein sequences from, higher plants, mosses and liverworts, different groups of eukaryotic algae and cyanobacteria. CP12 was identified throughout the Plantae, apart from in the Prasinophyceae. Within the Chromalveolata, two putative CP12 proteins have been found in the genomes of the diatom Thalassiosira pseudonana and the haptophyte Emiliania huxleyi, but specific searches in further chromalveolate genomes or EST datasets did not reveal any CP12 sequences in other Prymnesiophyceae, Dinophyceae or Pelagophyceae. A species from the Euglenophyceae within the Excavata also appeared to lack CP12. Phylogenetic analysis showed a clear separation into a number of higher taxonomic clades and among different forms of CP12 in higher plants. Cyanobacteria, Chlorophyceae, Rhodophyta and Glaucophyceae, Bryophyta, and the CP12-3 forms in higher plants all form separate clades. The degree of disorder of CP12 was higher in higher plants than in the eukaryotic algae and cyanobacteria apart from the green algal class Mesostigmatophyceae, which is ancestral to the streptophytes. This suggests that CP12 has evolved to become more flexible and possibly take on more general roles. Different features of the CP12 sequences in the different taxonomic groups and their potential functions and interactions in the Calvin cycle are discussed.  相似文献   

12.
Diatoms form large spring blooms in lakes and oceans, providing fuel for higher trophic levels at the start of the growing season. Some of the diatom blooms, however, are not grazed by filter-feeding zooplankton like Daphnia due to their large size. Several of these large diatoms are susceptible to chytrid infections. Zoospores of chytrids appeared to be excellent food for Daphnia, both in terms of size, shape, and quality (PUFAs and cholesterol). Thus, zoospores of chytrids can bridge the gap between inedible diatoms and Daphnia. In order to examine the effects of diatoms and chytrids on the survival of copepods, we performed one grazing and one survival experiment. The grazing experiment revealed that the diatom, Asterionella formosa, was not grazed by the copepod, Eudiaptomus gracilis, even after being infected by the chytrid Zygorhizidium planktonicum. However, carbon and nitrogen concentrations were significantly reduced by E. gracilis only when A. formosa was infected by Z. planktonicum, indicating that the chytrids might facilitate material transfer from inedible diatoms to the copepods. The survival experiment revealed that E. gracilis lived shorter with A. formosa than with the cryptophyta Cryptomonas pyrenoidifera. However, the survival of E. gracilis increased significantly in the treatment where A. formosa cells were infected by Z. planktonicum. Since E. gracilis could not graze A. formosa cells due to their large colonial forms, E. gracilis may acquire nutrients by grazing on the zoospores, and were so able to survive in the presence of the A. formosa. This provides new insights into the role of parasitic fungi in aquatic food webs, where chytrids may improve copepod survival during diatom blooms.  相似文献   

13.
Regulation of the Calvin–Benson cycle under varying light/dark conditions is a common property of oxygenic photosynthetic organisms and photosynthetic glyceraldehyde-3-phosphate dehydrogenase (GAPDH) is one of the targets of this complex regulatory system. In cyanobacteria and most algae, photosynthetic GAPDH is a homotetramer of GapA subunits which do not contain regulatory domains. In these organisms, dark-inhibition of the Calvin–Benson cycle involves the formation of a kinetically inhibited supramolecular complex between GAPDH, the regulatory peptide CP12 and phosphoribulokinase. Conditions prevailing in the dark, i.e. oxidation of thioredoxins and low NADP(H)/NAD(H) ratio promote aggregation. Although this regulatory system has been inherited in higher plants, these phototrophs contain in addition a second type of GAPDH subunits (GapB) resulting from the fusion of GapA with the C-terminal half of CP12. Heterotetrameric A2B2-GAPDH constitutes the major photosynthetic GAPDH isoform of higher plants chloroplasts and coexists with CP12 and A4-GAPDH. GapB subunits of A2B2-GAPDH have inherited from CP12 a regulatory domain (CTE for C-terminal extension) which makes the enzyme sensitive to thioredoxins and pyridine nucleotides, resembling the GAPDH/CP12/PRK system. The two systems are similar in other respects: oxidizing conditions and low NADP(H)/NAD(H) ratios promote aggregation of A2B2-GAPDH into strongly inactivated A8B8-GAPDH hexadecamers, and both CP12 and CTE specifically affect the NADPH-dependent activity of GAPDH. The alternative, lower activity with NADH is always unaffected. Based on the crystal structure of spinach A4-GAPDH and the analysis of site-specific mutants, a model of the autonomous (CP12-independent) regulatory mechanism of A2B2-GAPDH is proposed. Both CP12 and CTE seem to regulate different photosynthetic GAPDH isoforms according to a common and ancient molecular mechanism.  相似文献   

14.
15.
16.
In Chlamydomonas reinhardtii, glyceraldehyde-3-phosphate dehydrogenase (GAPDH) consists of four GapA subunits. This A4 GAPDH is not autonomously regulated, as the regulatory cysteine residues present on GapB subunits are missing in GapA subunits. The regulation of A4 GAPDH is provided by another protein, CP12. To determine the molecular mechanisms of regulation of A4 GAPDH, we mutated three residues (R82, R190, and S195) of GAPDH of C. reinhardtii. Kinetic studies of GAPDH mutants showed the importance of residue R82 in the specificity of GAPDH for NADPH, as previously shown for the spinach enzyme. The cofactor NADPH was not stabilized through the 2'-phosphate by the serine 195 residue of the algal GAPDH, unlike the case in spinach. The mutation of R190 also led to a structural change that was not observed in the spinach enzyme. This mutation led to a loss of activity for NADPH and NADH, indicating the crucial role of this residue in maintaining the algal GAPDH structure. Finally, the interaction between GAPDH mutants and wild-type and mutated CP12 was analyzed by immunoblotting experiments, surface plasmon resonance, and kinetic studies. The results obtained with these approaches highlight the involvement of the last residue of CP12, Asp80, in modulating the activity of GAPDH by preventing access of the cofactor NADPH to the active site. These results help us to bridge the gap between our knowledge of structure and our understanding of functional biology in GAPDH regulation.  相似文献   

17.
Eukaryotic RUBISCO appears in two sequence‐diverging forms, known as red‐like (present in nongreen algae) and green‐like (of green algae and higher plants) types. Oxidation of cysteines from green‐like RUBISCOs is known to result in conformational changes that inactivate the enzyme and render a relaxed structure more prone to proteolytic attack. These changes may have regulatory value for green algae and higher plants, promoting RUBISCO catabolism under stress conditions. We compare here red‐like RUBISCOs from several diatoms with a representative green‐like RUBISCO from Chlamydomonas reinhardtii, paying special attention to the cysteine‐dependent redox properties. Purified diatom RUBISCO preparations displayed a specific carboxylase activity about one order of magnitude lower than that of the C. reinhardtii P. A. Dang. enzyme. Despite having different patterns of cysteine residues in their primary sequence, the red‐like enzymes from diatoms inactivated also through oxidation of cysteine sulfhydryls to disulfides with a transition midpoint identical to that of the green‐like forms. Cysteine oxidation resulted also in structural modifications of the diatom RUBISCOs, as recognized by a higher sensitivity of the oxidized enzyme to in vitro proteolysis. The coincident redox properties of red‐ and green‐like RUBISCO types suggest that these changes are part of a physiologically significant regulatory mechanism that has been convergently implemented in both groups with a different set of cysteine residues.  相似文献   

18.
The structure of intertidal benthic diatoms assemblages in the Tagus estuary was investigated during a 2‐year survey, carried out in six stations with different sediment texture. Nonparametric multivariate analyses were used to characterize spatial and temporal patterns of the assemblages and to link them to the measured environmental variables. In addition, diversity and other features related to community physiognomy, such as size‐class or life‐form distributions, were used to describe the diatom assemblages. A total of 183 diatom taxa were identified during cell counts and their biovolume was determined. Differences between stations (analysis of similarity (ANOSIM), = 0.932) were more evident than temporal patterns (= 0.308) and mud content alone was the environmental variable most correlated to the biotic data (BEST, ρ = 0.863). Mudflat stations were typically colonized by low diversity diatom assemblages (H′ ~ 1.9), mainly composed of medium‐sized motile epipelic species (250–1,000 μm3), that showed species‐specific seasonal blooms (e.g., Navicula gregaria Donkin). Sandy stations had more complex and diverse diatom assemblages (H′ ~ 3.2). They were mostly composed by a large set of minute epipsammic species (<250 μm3) that, generally, did not show temporal patterns. The structure of intertidal diatom assemblages was largely defined by the interplay between epipelon and epipsammon, and its diversity was explained within the framework of the Intermediate Disturbance Hypothesis. However, the spatial distribution of epipelic and epipsammic life‐forms showed that the definition of both functional groups should not be over‐simplified.  相似文献   

19.
In darkened leaves the Calvin cycle enzymes glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and phosphoribulokinase (PRK) form a regulatory multi-enzyme complex with the small chloroplast protein CP12. GAPDH also forms a high molecular weight regulatory mono-enzyme complex. Given that there are different reports as to the number and subunit composition of these complexes and that enzyme regulatory mechanisms are known to vary between species, it was reasoned that protein-protein interactions may also vary between species. Here, this variation is investigated. This study shows that two different tetramers of GAPDH (an A2B2 heterotetramer and an A4 homotetramer) have the capacity to form part of the PRK/GAPDH/CP12 complex. The role of the PRK/GAPDH/CP12 complex is not simply to regulate the 'non-regulatory' A4 GAPDH tetramer. This study also demonstrates that the abundance and nature of PRK/GAPDH/CP12 interactions are not equal in all species and that whilst NAD enhances complex formation in some species, this is not sufficient for complex formation in others. Furthermore, it is shown that the GAPDH mono-enzyme complex is more abundant as a 2(A2B2) complex, rather than the larger 4(A2B2) complex. This smaller complex is sensitive to cellular metabolites indicating that it is an important regulatory isoform of GAPDH. This comparative study has highlighted considerable heterogeneity in PRK and GAPDH protein interactions between closely related species and the possible underlying physiological basis for this is discussed.  相似文献   

20.
Addition of the plant hormone 24-epibrassinolide to culture media stimulated the growth of a freshwater diatom, Asterionella formosa. The hormone stimulated activity of glyceraldehyde-3-phosphate dehydrogenase (GAPDH), a key enzyme from Calvin cycle, by 6-fold. Other key metabolic enzymes, phosphofructokinase and malate dehydrogenase were also stimulated but to a lesser extent. The activity of glucose-6-phosphate dehydrogenase, involved in the oxidative pentose phosphate pathway, also increased in the presence of the hormone but only under non reducing conditions. In cells stimulated by epibrassinolide, activated enzymes were sensitive to oxidized-DTT. GAPDH purified from cells grown in the presence of the hormone was not associated with a small protein of 8.5 kDa shown to be similar to CP12. Consequently the activity of GAPDH was no longer regulated by either oxidizing or reducing conditions. Among enzymes that, like GAPDH, responded positively to reducing agent were fructose-1,6-bisphosphatase (FBPase) and glucose-6-phosphate dehydrogenase (G6PDH). These enzymes were also sensitive to, and were negatively regulated by, oxidized-DTT. The activities in extracts from illuminated cells differed from those from darkened cells: FBPase, G6PDH and GAPDH, that were activated by DTT in darkened cells were no more activated in illuminated cells, but were oxidized by oxidized-DTT. Thus, oxidizing or reducing conditions mimic the conditions in dark and light, respectively. Unlike the other enzymes, phosphofructokinase (PFK) was inhibited by DTT but oxidized-DTT reversed this effect. The enzymes shown to be redox regulated in vitro by reduction/oxidation are very likely candidates for regulation in vivo by thioredoxins.  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