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.
Ribulose‐1,5‐bisphosphate carboxylase/oxygenase (Rubisco) plays a central role in carbon dioxide fixation on our planet. Rubisco from a hyperthermophilic archaeon Thermococcus kodakarensis (Tk‐Rubisco) shows approximately twenty times the activity of spinach Rubisco at high temperature, but only one‐eighth the activity at ambient temperature. We have tried to improve the activity of Tk‐Rubisco at ambient temperature, and have successfully constructed several mutants which showed higher activities than the wild‐type enzyme both in vitro and in vivo. Here, we designed new Tk‐Rubisco mutants based on its three‐dimensional structure and a sequence comparison of thermophilic and mesophilic plant Rubiscos. Four mutations were introduced to generate new mutants based on this strategy, and one of the four mutants, T289D, showed significantly improved activity compared to that of the wild‐type enzyme. The crystal structure of the Tk‐Rubisco T289D mutant suggested that the increase in activity was due to mechanisms distinct from those involved in the improvement in activity of Tk‐Rubisco SP8, a mutant protein previously reported to show the highest activity at ambient temperature. Combining the mutations of T289D and SP8 successfully generated a mutant protein (SP8‐T289D) with the highest activity to date both in vitro and in vivo. The improvement was particularly pronounced for the in vivo activity of SP8‐T289D when introduced into the mesophilic, photosynthetic bacterium Rhodopseudomonas palustris, which resulted in a strain with nearly two‐fold higher specific growth rates compared to that of a strain harboring the wild‐type enzyme at ambient temperature. Proteins 2016; 84:1339–1346. © 2016 Wiley Periodicals, Inc.  相似文献   

2.
The hyperthermophilic archaeon Thermococcus kodakaraensis harbors a type III ribulose 1,5-bisphosphate carboxylase/oxygenase (Rbc(Tk)). It has previously been shown that Rbc(Tk) is capable of supporting photoautotrophic and photoheterotrophic growth in a mesophilic host cell, Rhodopseudomonas palustris Delta3, whose three native Rubisco genes had been disrupted. Here, we have examined the enzymatic properties of Rbc(Tk) at 25 degrees C and have constructed mutant proteins in order to enhance its performance in mesophilic host cells. Initial sites for mutagenesis were selected by focusing on sequence differences in the loop 6 and alpha-helix 6 regions among Rbc(Tk) and the enzymes from spinach (mutant proteins SP1 to SP7), Galdieria partita (GP1 and GP2), and Rhodospirillum rubrum (RR1). Loop 6 of Rbc(Tk) is one residue longer than those found in the spinach and G. partita enzymes, and replacing Rbc(Tk) loop 6 with these regions led to dramatic decreases in activity. Six mutant enzymes retaining significant levels of Rubisco activity were selected, and their genes were introduced into R. palustris Delta3. Cells harboring mutant protein SP6 displayed a 31% increase in the specific growth rate under photoheterotrophic conditions compared to cells harboring wild-type Rbc(Tk). SP6 corresponds to a complete substitution of the original alpha-helix 6 of Rbc(Tk) with that of the spinach enzyme. Compared to wild-type Rbc(Tk), the purified SP6 mutant protein exhibited a 30% increase in turnover number (k(cat)) of the carboxylase activity and a 17% increase in the k(cat)/K(m) value. Based on these results, seven further mutant proteins were designed and examined. The results confirmed the importance of the length of loop 6 in Rbc(Tk) and also led to the identification of specific residue changes that resulted in an increase in the turnover number of Rbc(Tk) at ambient temperatures.  相似文献   

3.
The pyrenoid is a proteinaceous structure found in the chloroplast of most unicellular algae. Various studies indicate that ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) is present in the pyrenoid, although the fraction of Rubisco localized there remains controversial. Estimates of the amount of Rubisco in the pyrenoid of Chlamydomonas reinhardtii range from 5% to nearly 100%. Using immunolocalization, the amount of Rubisco localized to the pyrenoid or to the chloroplast stroma was estimated for C. reinhardtii cells grown under different conditions. It was observed that the amount of Rubisco in the pyrenoid varied with growth condition; about 40% was in the pyrenoid when the cells were grown under elevated CO2 and about 90% with ambient CO2. In addition, it is likely that pyrenoidal Rubisco is active in CO2 fixation because in vitro activity measurements showed that most of the Rubisco must be active to account for CO2-fixation rates observed in whole cells. These results are consistent with the idea that the pyrenoid is the site of CO2 fixation in C. reinhardtii and other unicellular algae containing CO2-concentrating mechanisms.  相似文献   

4.
The Calvin-Benson-Bassham cycle is responsible for carbon dioxide fixation in all plants, algae, and cyanobacteria. The enzyme that catalyzes the carbon dioxide-fixing reaction is ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). Rubisco from a hyperthermophilic archaeon Thermococcus kodakarensis (Tk-Rubisco) belongs to the type III group, and shows high activity at high temperatures. We have previously found that replacement of the entire α-helix 6 of Tk-Rubisco with the corresponding region of the spinach enzyme (SP6 mutant) results in an improvement of catalytic performance at mesophilic temperatures, both in vivo and in vitro, whereas the former and latter half-replacements of the α-helix 6 (SP4 and SP5 mutants) do not yield such improvement. We report here the crystal structures of the wild-type Tk-Rubisco and the mutants SP4 and SP6, and discuss the relationships between their structures and enzymatic activities. A comparison among these structures shows the movement and the increase of temperature factors of α-helix 6 induced by four essential factors. We thus supposed that an increase in the flexibility of the α-helix 6 and loop 6 regions was important to increase the catalytic activity of Tk-Rubisco at ambient temperatures. Based on this structural information, we constructed a new mutant, SP5-V330T, which was designed to have significantly greater flexibility in the above region, and it proved to exhibit the highest activity among all mutants examined to date. The thermostability of the SP5-V330T mutant was lower than that of wild-type Tk-Rubisco, providing further support on the relationship between flexibility and activity at ambient temperatures.  相似文献   

5.
The presence or absence of molecular oxygen has been shown to play a crucial role in the degradability of haloaromatic compounds. In the present study, it was shown that anaerobic phototrophic 3-chlorobenzoate (3CBA) metabolism by Rhodopseudomonas palustris DCP3 is oxygen tolerant up to a concentration of 3 μM O2. Simultaneous oxidation of an additional carbon source permitted light-dependent anaerobic 3CBA degradation at oxygen input levels which, in the absence of such an additional compound, would result in inhibition of light-dependent dehalogenation. Experiments under the same experimental conditions with strain DCP3 in coculture with an aerobic 3CBA-utilizing heterotroph, Alcaligenes sp. strain L6, revealed that light-dependent dehalogenation of 3CBA did not occur. Under both oxygen limitation (O2 < 0.1 μM) and low oxygen concentrations (3 μM O2), all the 3CBA was metabolized by the aerobic heterotroph. These data suggest that biodegradation of (halo)aromatics by photoheterotrophic bacteria such as R. palustris DCP3 may be restricted to anoxic photic environments.  相似文献   

6.
We report for the first time an analysis of the ATPase activity of human DNA topoisomerase (topo) IIβ. We show that topo IIβ is a DNA-dependent ATPase that appears to fit Michaelis–Menten kinetics. The ATPase activity is stimulated 44-fold by DNA. The kcat for ATP hydrolysis by human DNA topo IIβ in the presence of DNA is 2.25 s–1. We have characterised a topo IIβ derivative which carries a mutation in the ATPase domain (S165R). S165R reduced the kcat for ATP hydrolysis by 7-fold, to 0.32 s–1, while not significantly altering the apparent Km. The specificity constant for the interaction between ATP and topo IIβ (kcat/Kmapp) showed a 90% reduction for βS165R. The DNA binding affinity and ATP-independent DNA cleavage activity of the enzyme are unaffected by this mutation. However, the strand passage activity is reduced by 80%, presumably due to reduced ATP hydrolysis. The mutant enzyme is unable to complement ts yeast topo II in vivo. We have used computer modelling to predict the arrangement of key residues at the ATPase active site of topo IIβ. Ser165 is predicted to lie very close to the bound nucleotide, and the S165R mutation could thus influence both ATP binding and ADP dissociation.  相似文献   

7.
Purple nonsulfur bacteria grow photoheterotrophically by using light for energy and organic compounds for carbon and electrons. Disrupting the activity of the CO2-fixing Calvin cycle enzyme, ribulose 1,5-bisphosphate carboxylase (RubisCO), prevents photoheterotrophic growth unless an electron acceptor is provided or if cells can dispose of electrons as H2. Such observations led to the long-standing model wherein the Calvin cycle is necessary during photoheterotrophic growth to maintain a pool of oxidized electron carriers. This model was recently challenged with an alternative model wherein disrupting RubisCO activity prevents photoheterotrophic growth due to the accumulation of toxic ribulose-1,5-bisphosphate (RuBP) (D. Wang, Y. Zhang, E. L. Pohlmann, J. Li, and G. P. Roberts, J. Bacteriol. 193:3293-3303, 2011, http://dx.doi.org/10.1128/JB.00265-11). Here, we confirm that RuBP accumulation can impede the growth of Rhodospirillum rubrum (Rs. rubrum) and Rhodopseudomonas palustris (Rp. palustris) RubisCO-deficient (ΔRubisCO) mutants under conditions where electron carrier oxidation is coupled to H2 production. However, we also demonstrate that Rs. rubrum and Rp. palustris Calvin cycle phosphoribulokinase mutants that cannot produce RuBP cannot grow photoheterotrophically on succinate unless an electron acceptor is provided or H2 production is permitted. Thus, the Calvin cycle is still needed to oxidize electron carriers even in the absence of toxic RuBP. Surprisingly, Calvin cycle mutants of Rs. rubrum, but not of Rp. palustris, grew photoheterotrophically on malate without electron acceptors or H2 production. The mechanism by which Rs. rubrum grows under these conditions remains to be elucidated.  相似文献   

8.
The act gene of Variovorax paradoxus TBEA6 encodes a succinyl-CoA:3-sulfinopropionate coenzyme A (CoA)-transferase, ActTBEA6 (2.8.3.x), which catalyzes the activation of 3-sulfinopropionate (3SP), an intermediate during 3,3′-thiodipropionate (TDP) degradation. In a previous study, accumulation of 3SP was observed in a Tn5::mob-induced mutant defective in growth on TDP. In contrast to the wild type and all other obtained mutants, this mutant showed no growth when 3SP was applied as the sole source of carbon and energy. The transposon Tn5::mob was inserted in a gene showing high homology to class III CoA-transferases. In the present study, analyses of the translation product clearly allocated ActTBEA6 to this protein family. The predicted secondary structure indicates the lack of a C-terminal α-helix. ActTBEA6 was heterologously expressed in Escherichia coli Lemo21(DE3) and was then purified by Ni-nitrilotriacetic acid (NTA) affinity chromatography. Analytical size exclusion chromatography revealed a homodimeric structure with a molecular mass of 96 ± 3 kDa. Enzyme assays identified succinyl-CoA, itaconyl-CoA, and glutaryl-CoA as potential CoA donors and unequivocally verified the conversion of 3SP to 3SP-CoA. Kinetic studies revealed an apparent Vmax of 44.6 μmol min−1 mg−1 for succinyl-CoA, which corresponds to a turnover number of 36.0 s−1 per subunit of ActTBEA6. For 3SP, the apparent Vmax was determined as 46.8 μmol min−1 mg−1, which corresponds to a turnover number of 37.7 s−1 per subunit of ActTBEA6. The apparent Km values were 0.08 mM for succinyl-CoA and 5.9 mM for 3SP. Nonetheless, the V. paradoxus Δact mutant did not reproduce the phenotype of the Tn5::mob-induced mutant. This defined deletion mutant was able to utilize TDP or 3SP as the sole carbon source, like the wild type. Complementation of the Tn5::mob-induced mutant with pBBR1MCS5::acdDPN7 partially restored growth on 3SP, which indicated a polar effect of the Tn5::mob transposon on acdTBEA6, located downstream of actTBEA6.  相似文献   

9.
There has been much interest in the chloroplast-encoded large subunit of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) as a target for engineering an increase in net CO2 fixation in photosynthesis. Improvements in the enzyme would lead to an increase in the production of food, fiber, and renewable energy. Although the large subunit contains the active site, a family of rbcS nuclear genes encodes the Rubisco small subunits, which can also influence the carboxylation catalytic efficiency and CO2/O2 specificity of the enzyme. To further define the role of the small subunit in Rubisco function, small subunits from spinach, Arabidopsis, and sunflower were assembled with algal large subunits by transformation of a Chlamydomonas reinhardtii mutant that lacks the rbcS gene family. Foreign rbcS cDNAs were successfully expressed in Chlamydomonas by fusing them to a Chlamydomonas rbcS transit peptide sequence engineered to contain rbcS introns. Although plant Rubisco generally has greater CO2/O2 specificity but a lower carboxylation Vmax than Chlamydomonas Rubisco, the hybrid enzymes have 3–11% increases in CO2/O2 specificity and retain near normal Vmax values. Thus, small subunits may make a significant contribution to the overall catalytic performance of Rubisco. Despite having normal amounts of catalytically proficient Rubisco, the hybrid mutant strains display reduced levels of photosynthetic growth and lack chloroplast pyrenoids. It appears that small subunits contain the structural elements responsible for targeting Rubisco to the algal pyrenoid, which is the site where CO2 is concentrated for optimal photosynthesis.  相似文献   

10.
The plasmid pGKE75-catA138T, which comprises pUC18 and the catA138T gene encoding thermostable chloramphenicol acetyltransferase with an A138T amino acid replacement (CATA138T), serves as an Escherichia coli-Geobacillus kaustophilus shuttle plasmid that confers moderate chloramphenicol resistance on G. kaustophilus HTA426. The present study examined the thermoadaptation-directed mutagenesis of pGKE75-catA138T in an error-prone thermophile, generating the mutant plasmid pGKE75αβ-catA138T responsible for substantial chloramphenicol resistance at 65°C. pGKE75αβ-catA138T contained no mutation in the catA138T gene but had two mutations in the pUC replicon, even though the replicon has no apparent role in G. kaustophilus. Biochemical characterization suggested that the efficient chloramphenicol resistance conferred by pGKE75αβ-catA138T is attributable to increases in intracellular CATA138T and acetyl-coenzyme A following a decrease in incomplete forms of pGKE75αβ-catA138T. The decrease in incomplete plasmids may be due to optimization of plasmid replication by RNA species transcribed from the mutant pUC replicon, which were actually produced in G. kaustophilus. It is noteworthy that G. kaustophilus was transformed with pGKE75αβ-catA138T using chloramphenicol selection at 60°C. In addition, a pUC18 derivative with the two mutations propagated in E. coli at a high copy number independently of the culture temperature and high plasmid stability. Since these properties have not been observed in known plasmids, the outcomes extend the genetic toolboxes for G. kaustophilus and E. coli.  相似文献   

11.
To ascertain whether position 131 of a mesophilic protease, subtilisin BPN′, is a potential critical site for cold adaptation as screened by evolutionary engineering (S. Taguchi, A. Ozaki, and H. Momose, Appl. Environ. Microbiol. 64:492–495, 1998), a full set of subtilisin BPN′ mutants with mutations at position 131 was constructed by site-saturation mutagenesis. All mutated enzymes were measured for specific activity at 10°C by the quantitative titer microplate assay system using polyclonal antibody against subtilisin BPN′ and a synthetic chromogenic substrate. All the mutants exhibited proteolytic activities almost the same as or higher than that of the wild-type enzyme, suggesting that position 131 may be important for cold adaptation. In comparison with the wild type, purified mutants G131F, G131R, G131M, and G131W were found to acquire proteolytic activities (kcat/Km) at 10°C that were 150, 94, 84, and 50% higher, respectively. In particular, for the G131F mutant, temperature dependency in enzyme activity was shown by an increase in kcat and a decrease in Km. All of these amino acid substitution mutants, G131F, G131R, G131M, and G131W, acquired increased proteolytic activities at 10°C for three different synthetic peptide substrates but no increase in caseinolytic activity. Furthermore, they all conferred thermolability on the enzyme to differing extents in terms of the half-life of enzyme inactivation at 60°C. No significant correlation was found between the amino acids preferred for cold adaptation surveyed here and those present at position 131 of subtilisin of psychrophilic cells naturally occurring in cold environments. Based on these findings, position 131 is a contributor in artificial evolution for acquiring a cold-active character and may not be related to physiological requirements for subtilisin-producing cells living in cold environments. Therefore, saturation mutagenesis would be effective in achieving rapid improvement in protein properties via evolutionary engineering.  相似文献   

12.
Microbial communities in a shallow submarine hydrothermal system near Taketomi Island, Japan, were investigated using cultivation-based and molecular techniques. The main hydrothermal activity occurred in a craterlike basin (depth, ~23 m) on the coral reef seafloor. The vent fluid (maximum temperature, >52°C) contained 175 μM H2S and gas bubbles mainly composed of CH4 (69%) and N2 (29%). A liquid serial dilution cultivation technique targeting a variety of metabolism types quantified each population in the vent fluid and in a white microbial mat located near the vent. The most abundant microorganisms cultivated from both the fluid and the mat were autotrophic sulfur oxidizers, including mesophilic Thiomicrospira spp. and thermophilic Sulfurivirga caldicuralii. Methane oxidizers were the second most abundant organisms in the fluid; one novel type I methanotroph exhibited optimum growth at 37°C, and another novel type I methanotroph exhibited optimum growth at 45°C. The number of hydrogen oxidizers cultivated only from the mat was less than the number of sulfur and methane oxidizers, although a novel mesophilic hydrogen-oxidizing member of the Epsilonproteobacteria was isolated. Various mesophilic to hyperthermophilic heterotrophs, including sulfate-reducing Desulfovibrio spp., iron-reducing Deferribacter sp., and sulfur-reducing Thermococcus spp., were also cultivated. Culture-independent 16S rRNA gene clone analysis of the vent fluid and mat revealed highly diverse archaeal communities. In the bacterial community, S. caldicuralii was identified as the predominant phylotype in the fluid (clonal frequency, 25%). Both bacterial clone libraries indicated that there were bacterial communities involved in sulfur, hydrogen, and methane oxidation and sulfate reduction. Our results indicate that there are unique microbial communities that are sustained by active chemosynthetic primary production rather than by photosynthetic production in a shallow hydrothermal system where sunlight is abundant.  相似文献   

13.
The hyperthermophilic archaeon, Thermococcus kodakaraensis KOD1 harbors a structurally novel, Type III Rubisco (Rbc(Tk)). In terms of protein engineering of Rubiscos, the enzyme may provide an alternative target to the conventional Type I and Type II enzymes. With a future aim to improve the catalytic properties of Rbc(Tk), here we examined whether or not the enzyme could support growth of a mesophilic organism dependent on CO2 fixation. Via double-crossover homologous recombination, we first deleted three Rubisco genes present on the chromosome of the photosynthetic mesophile Rhodopseudomonas palustris No. 7. The mutant strain (delta3) could neither grow under photoautotrophic nor photoheterotrophic conditions. We introduced the rbc(Tk) gene into strain delta3 either on a plasmid, or by integrating the gene onto the chromosome. The two transformant strains harboring rbc(Tk) displayed growth under photoautotrophic and photoheterotrophic conditions, both dependent on CO2 fixation. Specific growth rates and Rubisco activity levels were compared under photoheterotrophic conditions among the two transformants and the wild-type strain. We observed that the levels of Rubisco activity in the respective cell-free extracts correlated well with the specific growth rates. Immunoprecipitation experiments revealed that Rubisco activity detected in the transformants was derived solely from Rbc(Tk). These results demonstrated that the Type III Rbc(Tk) from a hyperthermophile could support CO2 fixation in a mesophilic organism, and that the specific growth rate of the transformant can be used as a convenient parameter for selection of engineered proteins with improved Rubisco activity.  相似文献   

14.
The activation of antithrombin (AT) by heparin facilitates the exosite-dependent interaction of the serpin with factors IXa (FIXa) and Xa (FXa), thereby improving the rate of reactions by 300- to 500-fold. Relative to FXa, AT inhibits FIXa with ∼40-fold slower rate constant. Structural data suggest that differences in the residues of the 39-loop (residues 31–41) may partly be responsible for the differential reactivity of the two proteases with AT. This loop is highly acidic in FXa, containing three Glu residues at positions 36, 37, and 39. By contrast, the loop is shorter by one residue in FIXa (residue 37 is missing), and it contains a Lys and an Asp at positions 36 and 39, respectively. To determine whether differences in the residues of this loop contribute to the slower reactivity of FIXa with AT, we prepared an FIXa/FXa chimera in which the 39-loop of the protease was replaced with the corresponding loop of FXa. The chimeric mutant cleaved a FIXa-specific chromogenic substrate with normal catalytic efficiency, however, the mutant exhibited ∼5-fold enhanced reactivity with AT specifically in the absence of the cofactor, heparin. Further studies revealed that the FIXa mutant activates factor X with ∼4-fold decreased kcat and ∼2-fold decreased Km, although the mutant interacted normally with factor VIIIa. Based on these results we conclude that residues of the 39-loop regulate the cofactor-independent interaction of FIXa with its physiological inhibitor AT and substrate factor X.  相似文献   

15.
The first x-ray crystal structure has been solved for an activated transition-state analog-bound form II ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). This enzyme, from Rhodopseudomonas palustris, assembles as a unique hexamer with three pairs of catalytic large subunit homodimers around a central 3-fold symmetry axis. This oligomer arrangement is unique among all known Rubisco structures, including the form II homolog from Rhodospirillum rubrum. The presence of a transition-state analog in the active site locked the activated enzyme in a “closed” conformation and revealed the positions of critical active site residues during catalysis. Functional roles of two form II-specific residues (Ile165 and Met331) near the active site were examined via site-directed mutagenesis. Substitutions at these residues affect function but not the ability of the enzyme to assemble. Random mutagenesis and suppressor selection in a Rubisco deletion strain of Rhodobacter capsulatus identified a residue in the amino terminus of one subunit (Ala47) that compensated for a negative change near the active site of a neighboring subunit. In addition, substitution of the native carboxyl-terminal sequence with the last few dissimilar residues from the related R. rubrum homolog increased the enzyme''s kcat for carboxylation. However, replacement of a longer carboxyl-terminal sequence with termini from either a form III or a form I enzyme, which varied both in length and sequence, resulted in complete loss of function. From these studies, it is evident that a number of subtle interactions near the active site and the carboxyl terminus account for functional differences between the different forms of Rubiscos found in nature.  相似文献   

16.
Heuer B  Portis AR 《Plant physiology》1990,93(4):1511-1513
Optimal storage conditions to retain ribulose 1, 5-bisphosphate carboxylase/oxygenase (Rubisco) activity were investigated. The soluble spinach (Spinacia oleracea) enzyme was pretreated with its activators, Mg2+ and HCO3, and then stored for up to 30 days at 4 or −18°C or in liquid N2. Cold inactivation and conformational changes were suggested to be involved during Rubisco storage in the cold, leading to its inactivation. Pretreatment of the enzyme with Mg2+ and CO2 and subsequent storage at either 4°C or in liquid N2 or flushing the samples with N2 and rapid freezing and storage in liquid N2 are recommended as storage procedures. These storage treatments will prevent inactivation, so that full original specific activity will be preserved.  相似文献   

17.
The terminal oxidase complexes encoded by coxMNOP and coxWXYZ were studied by analysis of mutations in each of the two oxidases. Carbon monoxide difference spectra obtained from membranes of coxMNOP mutant bacteroids were like those obtained for the wild type, whereas bacteroid membranes of a coxWXYZ mutant were deficient in CO-reactive cytochrome b. Experiments involving cyanide inhibition of oxidase activity were consistent with the conclusion that the coxX mutant is deficient in a membrane-associated O2-binding component. The viable cell number (bacteria that could be recultured from crushed nodules) was 20 to 29% lower for the coxX mutant than for the wild-type or the CoxN strain. In three separate greenhouse studies, nodules of a coxX mutant had significantly lower (28 to 34% less) acetylene reduction rates than the wild-type nodules did, and plants inoculated with a double mutant (coxMNOP coxWZYZ) had rates 30% lower than those of wild-type-inoculated plants.  相似文献   

18.
The regulation of Rubisco activity was investigated under high, constant photosynthetic photon flux density during the diurnal phases of Crassulacean acid metabolism in Kalanchoë daigremontiana Hamet et Perr. During phase I, a significant period of nocturnal, C4-mediated CO2 fixation was observed, with the generated malic acid being decarboxylated the following day (phase III). Two periods of daytime atmospheric CO2 fixation occurred at the beginning (phase II, C4–C3 carboxylation) and end (phase IV, C3–C4 carboxylation) of the day. During the 1st h of the photoperiod, when phosphoenolpyruvate carboxylase was still active, the highest rates of atmospheric CO2 uptake were observed, coincident with the lowest rates of electron transport and minimal Rubisco activity. Over the next 1 to 2 h of phase II, carbamylation increased rapidly during an initial period of decarboxylation. Maximal carbamylation (70%–80%) was reached 2 h into phase III and was maintained under conditions of elevated CO2 resulting from malic acid decarboxylation. Initial and total Rubisco activity increased throughout phase III, with maximal activity achieved 9 h into the photoperiod at the beginning of phase IV, as atmospheric CO2 uptake recommenced. We suggest that the increased enzyme activity supports assimilation under CO2-limited conditions at the start of phase IV. The data indicate that Rubisco activity is modulated in-line with intracellular CO2 supply during the daytime phases of Crassulacean acid metabolism.  相似文献   

19.
This work addresses the binding, cleavage and dissociation rates for the substrate and products of a synthetic RNaseA mimic that was combinatorially selected using chemically modified nucleoside triphosphates. This trans-cleaving DNAzyme, 925-11t, catalyzes sequence-specific ribophosphodiester hydrolysis in the total absence of a divalent metal cation, and in low ionic strength at pH 7.5 and in the presence of EDTA. It is the first such sequence capable of multiple turnover. 925-11t consists of 31 bases, 18 of which form a catalytic domain containing 4 imidazole and 6 allylamino modified nucleotides. This sequence cleaves the 15 nt long substrate, S1, at one embedded ribocytosine at the eighth position to give a 5′-product terminating in a 2′,3′-phosphodiester and a 3′-product terminating in a 5′-OH. Under single turnover conditions at 24°C, 925-11t displays a maximum first-order rate constant, kcat, of 0.037 min−1 and a catalytic efficiency, kcat/Km, of 5.3 × 105 M−1 min−1. The measured value of kcat under catalyst excess conditions agrees with the value of kcat observed for steady-state multiple turnover, implying that slow product release is not rate limiting with respect to multiple turnover. The substrate specificity of 925-11t was gauged in terms of kcat values for substrate sequence variants. Base substitutions on the scissile ribose and at the two bases immediately downstream decrease kcat values by a factor of 4 to 250, indicating that 925-11t displays significant sequence specificity despite the lack of an apparent Watson–Crick base-pairing scheme for recognition.  相似文献   

20.
A new method is presented for measurement of the CO2/O2 specificity factor of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco). The [14C]3-phosphoglycerate (PGA) from the Rubisco carboxylase reaction and its dilution by the Rubisco oxygenase reaction was monitored by directly measuring the specific radioactivity of PGA. 14CO2 fixation with Rubisco occurred under two reaction conditions: carboxylase with oxygenase with 40 micromolar CO2 in O2-saturated water and carboxylase only with 160 micromolar CO2 under N2. Detection of the specific radioactivity used the amount of PGA as obtained from the peak area, which was determined by pulsed amperometry following separation by high-performance anion exchange chromatography and the radioactive counts of the [14C]PGA in the same peak. The specificity factor of Rubisco from spinach (Spinacia oleracea L.) (93 ± 4), from the green alga Chlamydomonas reinhardtii (66 ± 1), and from the photosynthetic bacterium Rhodospirillum rubrum (13) were comparable with the published values measured by different methods.  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