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 15 毫秒
1.
Phosphorus is an essential component of macromolecules, like DNA, and central metabolic intermediates, such as sugar phosphates, and bacteria possess enzymes and control mechanisms that provide an optimal supply of phosphorus from the environment. UDP-sugar hydrolases and 5′ nucleotidases may play roles in signal transduction, as they do in mammals, in nucleotide salvage, as demonstrated for UshA of Escherichia coli, or in phosphorus metabolism. The Corynebacterium glutamicum gene ushA was found to encode a secreted enzyme which is active as a 5′ nucleotidase and a UDP-sugar hydrolase. This enzyme was synthesized and secreted into the medium when C. glutamicum was starved for inorganic phosphate. UshA was required for growth of C. glutamicum on AMP and UDP-glucose as sole sources of phosphorus. Thus, in contrast to UshA from E. coli, C. glutamicum UshA is an important component of the phosphate starvation response of this species and is necessary to access nucleotides and related compounds as sources of phosphorus.  相似文献   

2.
Fructooligosaccharides and their anhydrides are widely used as health-promoting foods and prebiotics. Various enzymes acting on β-D-fructofuranosyl linkages of natural fructan polymers have been used to produce functional compounds. However, enzymes that hydrolyze and form α-D-fructofuranosyl linkages have been less studied. Here, we identified the BBDE_2040 gene product from Bifidobacterium dentium (α-D-fructofuranosidase and difructose dianhydride I synthase/hydrolase from Bifidobacterium dentium [αFFase1]) as an enzyme with α-D-fructofuranosidase and α-D-arabinofuranosidase activities and an anomer-retaining manner. αFFase1 is not homologous with any known enzymes, suggesting that it is a member of a novel glycoside hydrolase family. When caramelized fructose sugar was incubated with αFFase1, conversions of β-D-Frup-(2→1)-α-D-Fruf to α-D-Fruf-1,2′:2,1′-β-D-Frup (diheterolevulosan II) and β-D-Fruf-(2→1)-α-D-Fruf (inulobiose) to α-D-Fruf-1,2′:2,1′-β-D-Fruf (difructose dianhydride I [DFA I]) were observed. The reaction equilibrium between inulobiose and DFA I was biased toward the latter (1:9) to promote the intramolecular dehydrating condensation reaction. Thus, we named this enzyme DFA I synthase/hydrolase. The crystal structures of αFFase1 in complex with β-D-Fruf and β-D-Araf were determined at the resolutions of up to 1.76 Å. Modeling of a DFA I molecule in the active site and mutational analysis also identified critical residues for catalysis and substrate binding. The hexameric structure of αFFase1 revealed the connection of the catalytic pocket to a large internal cavity via a channel. Molecular dynamics analysis implied stable binding of DFA I and inulobiose to the active site with surrounding water molecules. Taken together, these results establish DFA I synthase/hydrolase as a member of a new glycoside hydrolase family (GH172).  相似文献   

3.
Phloretin hydrolase catalyzes the hydrolytic C-C cleavage of phloretin to phloroglucinol and 3-(4-hydroxyphenyl)propionic acid during flavonoid degradation in Eubacterium ramulus. The gene encoding the enzyme was cloned by screening a gene library for hydrolase activity. The insert of a clone conferring phloretin hydrolase activity was sequenced. Sequence analysis revealed an open reading frame of 822 bp (phy), a putative promoter region, and a terminating stem-loop structure. The deduced amino acid sequence of phy showed similarities to a putative protein of the 2,4-diacetylphloroglucinol biosynthetic operon from Pseudomonas fluorescens. The phloretin hydrolase was heterologously expressed in Escherichia coli and purified. The molecular mass of the native enzyme was approximately 55 kDa as determined by gel filtration. The results of sodium dodecyl sulfate-polyacrylamide gel electrophoresis and the deduced amino acid sequence of phy indicated molecular masses of 30 and 30.8 kDa, respectively, suggesting that the enzyme is a homodimer. The recombinant phloretin hydrolase catalyzed the hydrolysis of phloretin to equimolar amounts of phloroglucinol and 3-(4-hydroxyphenyl)propionic acid. The optimal temperature and pH of the catalyzed reaction mixture were 37°C and 7.0, respectively. The Km for phloretin was 13 ± 3 μM and the kcat was 10 ± 2 s−1. The enzyme did not transform phloretin-2′-glucoside (phloridzin), neohesperidin dihydrochalcone, 1,3-diphenyl-1,3-propandione, or trans-1,3-diphenyl-2,3-epoxy-propan-1-one. The catalytic activity of the phloretin hydrolase was reduced by N-bromosuccinimide, o-phenanthroline, N-ethylmaleimide, and CuCl2 to 3, 20, 35, and 85%, respectively. Phloroglucinol and 3-(4-hydroxyphenyl)propionic acid reduced the activity to 54 and 70%, respectively.  相似文献   

4.
Sphingomonas paucimobilis SYK-6 has the ability to transform a lignin-related biphenyl compound, 2,2′-dihydroxy-3,3′-dimethoxy-5,5′-dicarboxybiphenyl (DDVA), to 5-carboxyvanillic acid (5CVA) via 2,2′,3-trihydroxy-3′-methoxy-5,5′-dicarboxybiphenyl (OH-DDVA). In the 4.9-kb HindIII fragment containing the OH-DDVA meta-cleavage dioxygenase gene (ligZ), we found a novel hydrolase gene (ligY) responsible for the conversion of the meta-cleavage compound of OH-DDVA to 5CVA. Incorporation of 18O from H218O into 5CVA indicated there was a hydrolytic conversion of the OH-DDVA meta-cleavage compound to 5CVA. LigY exhibited hydrolase activity only toward the meta-cleavage compound of OH-DDVA, suggesting its restricted substrate specificity.  相似文献   

5.
trans-2′-Carboxybenzalpyruvate hydratase-aldolase was purified from a phenanthrene-degrading bacterium, Nocardioides sp. strain KP7, and characterized. The purified enzyme was found to have molecular masses of 38 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and 113 kDa by gel filtration chromatography. Thus, the homotrimer of the 38-kDa subunit constituted an active enzyme. The Km and kcat values of this enzyme for trans-2′-carboxybenzalpyruvate were 50 μM and 13 s−1, respectively. trans-2′-Carboxybenzalpyruvate was transformed to 2-carboxybenzaldehyde and pyruvate by the action of this enzyme. The structural gene for this enzyme was cloned and sequenced; the length of this gene was 996 bp. The deduced amino acid sequence of this enzyme exhibited homology to those of trans-2′-hydroxybenzalpyruvate hydratase-aldolases from Pseudomonas putida PpG7 and Pseudomonas sp. strain C18.  相似文献   

6.
Thiocyanate hydrolase is a newly found enzyme from Thiobacillus thioparus THI 115 that converts thiocyanate to carbonyl sulfide and ammonia (Y. Katayama, Y. Narahara, Y. Inoue, F. Amano, T. Kanagawa, and H. Kuraishi, J. Biol. Chem. 267:9170–9175, 1992). We have cloned and sequenced the scn genes that encode the three subunits of the enzyme. The scnB, scnA, and scnC genes, arrayed in this order, contained open reading frames encoding sequences of 157, 126, and 243 amino acid residues, respectively, for the β, α, and γ subunits, respectively. Each open reading frame was preceded by a typical Shine-Dalgarno sequence. The deduced amino-terminal peptide sequences for the three subunits were in fair agreement with the chemically determined sequences. The protein molecular mass calculated for each subunit was compatible with that determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. From a computer analysis, thiocyanate hydrolase showed significant homologies to bacterial nitrile hydratases known to convert nitrile to the corresponding amide, which is further hydrolyzed by amidase to form acid and ammonia. The two enzymes were homologous over regions corresponding to almost the entire coding regions of the genes: the β and α subunits of thiocyanate hydrolase were homologous to the amino- and carboxyl-terminal halves of the β subunit of nitrile hydratase, and the γ subunit of thiocyanate hydrolase was homologous to the α subunit of nitrile hydratase. Comparisons of the catalytic properties of the two homologous enzymes support the model for the reaction steps of thiocyanate hydrolase that was previously presented on the basis of biochemical analyses.  相似文献   

7.
The solubility of orthophosphate (PO43−) in iron-rich sediments can be exceedingly low, limiting the bioavailability of this essential nutrient to microbial populations that catalyze critical biogeochemical reactions. Here we demonstrate that dissolved extracellular DNA can serve as a sole source of phosphorus, as well as carbon and energy, for metal-reducing bacteria of the genus Shewanella. Shewanella oneidensis MR-1, Shewanella putrefaciens CN32, and Shewanella sp. strain W3-18-1 all grew with DNA but displayed different growth rates. W3-18-1 exhibited the highest growth rate with DNA. While strain W3-18-1 displayed Ca2+-independent DNA utilization, both CN32 and MR-1 required millimolar concentrations of Ca2+ for growth with DNA. For S. oneidensis MR-1, the utilization of DNA as a sole source of phosphorus is linked to the activities of extracellular phosphatase(s) and a Ca2+-dependent nuclease(s), which are regulated by phosphorus availability. Mass spectrometry analysis of the extracellular proteome of MR-1 identified one putative endonuclease (SO1844), a predicted UshA (bifunctional UDP-sugar hydrolase/5′ nucleotidase), a predicted PhoX (calcium-activated alkaline phosphatase), and a predicted CpdB (bifunctional 2′,3′ cyclic nucleotide 2′ phosphodiesterase/3′ nucleotidase), all of which could play important roles in the extracellular degradation of DNA under phosphorus-limiting conditions. Overall, the results of this study suggest that the ability to use exogenous DNA as the sole source of phosphorus is widespread among the shewanellae, and perhaps among all prokaryotes, and may be especially important for nutrient cycling in metal-reducing environments.  相似文献   

8.
TrmJ proteins from the SPOUT methyltransferase superfamily are tRNA Xm32 modification enzymes that occur in bacteria and archaea. Unlike archaeal TrmJ, bacterial TrmJ require full-length tRNA molecules as substrates. It remains unknown how bacterial TrmJs recognize substrate tRNAs and specifically catalyze a 2′-O modification at ribose 32. Herein, we demonstrate that all six Escherichia coli (Ec) tRNAs with 2′-O-methylated nucleosides at position 32 are substrates of EcTrmJ, and we show that the elbow region of tRNA, but not the amino acid acceptor stem, is needed for the methylation reaction. Our crystallographic study reveals that full-length EcTrmJ forms an unusual dimer in the asymmetric unit, with both the catalytic SPOUT domain and C-terminal extension forming separate dimeric associations. Based on these findings, we used electrophoretic mobility shift assay, isothermal titration calorimetry and enzymatic methods to identify amino acids within EcTrmJ that are involved in tRNA binding. We found that tRNA recognition by EcTrmJ involves the cooperative influences of conserved residues from both the SPOUT and extensional domains, and that this process is regulated by the flexible hinge region that connects these two domains.  相似文献   

9.
Programmed RNA breakage is an emerging theme underlying cellular responses to stress, virus infection and defense against foreign species. In many cases, site-specific cleavage of the target RNA generates 2′,3′ cyclic phosphate and 5′-OH ends. For the damage to be repaired, both broken ends must be healed before they can be sealed by a ligase. Healing entails hydrolysis of the 2′,3′ cyclic phosphate to form a 3′-OH and phosphorylation of the 5′-OH to form a 5′-PO4. Here, we demonstrate that a polynucleotide kinase-phosphatase enzyme from Clostridium thermocellum (CthPnkp) can catalyze both of the end-healing steps of tRNA splicing in vitro. The route of tRNA repair by CthPnkp can be reprogrammed by a mutation in the 3′ end-healing domain (H189D) that yields a 2′-PO4 product instead of a 2′-OH. Whereas tRNA ends healed by wild-type CthPnkp are readily sealed by T4 RNA ligase 1, the H189D enzyme generates ends that are spliced by yeast tRNA ligase. Our findings suggest that RNA repair enzymes can evolve their specificities to suit a particular pathway.  相似文献   

10.
Biphenyl dioxygenase (BPH dox) oxidizes biphenyl on adjacent carbons to generate 2,3-dihydro-2,3-dihydroxybiphenyl in Comamonas testosteroni B-356 and in Pseudomonas sp. strain LB400. The enzyme comprises a two-subunit (α and β) iron sulfur protein (ISPBPH), a ferredoxin (FERBPH), and a ferredoxin reductase (REDBPH). B-356 BPH dox preferentially catalyzes the oxidation of the double-meta-substituted congener 3,3′-dichlorobiphenyl over the double-para-substituted congener 4,4′-dichlorobiphenyl or the double-ortho-substituted congener 2,2′-dichlorobiphenyl. LB400 BPH dox shows a preference for 2,2′-dichlorobiphenyl, and in addition, unlike B-356 BPH dox, it can catalyze the oxidation of selected chlorobiphenyls such as 2,2′,5,5′-tetrachlorobiphenyl on adjacent meta-para carbons. In this work, we examine the reactivity pattern of BPH dox toward various chlorobiphenyls and its capacity to catalyze the meta-para dioxygenation of chimeric enzymes obtained by exchanging the ISPBPH α or β subunit of strain B-356 for the corresponding subunit of strain LB400. These hybrid enzymes were purified by an affinity chromatography system as His-tagged proteins. Both types, the chimera with the α subunit of ISPBPH of strain LB400 and the β subunit of ISPBPH of strain B-356 (the αLB400βB-356 chimera) and the αB-356βLB400 chimera, were functional. Results with purified enzyme preparations showed for the first time that the ISPBPH β subunit influences BPH dox’s reactivity pattern toward chlorobiphenyls. Thus, if the α subunit were the sole determinant of the enzyme reactivity pattern, the αB-356βLB400 chimera should have behaved like B-356 ISPBPH; instead, its reactivity pattern toward the substrates tested was similar to that of LB400 ISPBPH. On the other hand, the αLB400βB-356 chimera showed features of both B-356 and LB400 ISPBPH where the enzyme was able to metabolize 2,2′- and 3,3′-dichlorobiphenyl and where it was able to catalyze the meta-para oxygenation of 2,2′,5,5′-tetrachlorobiphenyl.  相似文献   

11.
Ross C  Murray MG 《Plant physiology》1971,48(5):626-630
Mechanisms controlling conversion of orotic acid-6-14C to uridine-5′-phosphate in cotyledons of germinating Alaska peas (Pisum sativum L.) were investigated. The content of 5-phosphoribosyl-1-pyrophosphate was very low in dry seeds, increased to a maximum after about 12 hours of imbibition, and then rapidly declined. Orotidine-5′-phosphate pyrophosphorylase and orotidine-5′-phosphate decarboxylase activities more than doubled during the first 24 hours of germination and then also decreased. These results do not account for the continuous increases of orotate anabolism in such cotyledons as we observed previously. The initial increases in activities of these two enzymes were unaffected by cycloheximide, while the subsequent decreases were less rapid in the presence of this inhibitor. Activities of cotyledonary cytidine deaminase and uridine hydrolase also increased during imbibition, but the activity of only the latter showed a decrease after imbibition was completed. Cycloheximide inhibited the initial rapid increase in uridine hydrolase activity but had little effect on its subsequent decline. Cycloheximide had only slight inhibitory effects on the development of cytidine deaminase activity during the first 62 hours. The evidence suggests that uridine hydrolase might be synthesized de novo during the first few days of germination, but that the other three enzymes might not be.  相似文献   

12.
The intercellular distribution of assimilatory sulfate reduction enzymes between mesophyll and bundle sheath cells was analyzed in maize (Zea mays L.) and wheat (Triticum aestivum L.) leaves. In maize, a C4 plant, 96 to 100% of adenosine 5′-phosphosulfate sulfotransferase and 92 to 100% of ATP sulfurylase activity (EC 2.7.7.4) was detected in the bundle sheath cells. Sulfite reductase (EC 1.8.7.1) and O-acetyl-l-serine sulfhydrylase (EC 4.2.99.8) were found in both bundle sheath and mesophyll cell types. In wheat, a C3 species, ATP sulfurylase and adenosine 5′-phosphosulfate sulfotransferase were found at equivalent activities in both mesophyll and bundle sheath cells. Leaves of etiolated maize plants contained appreciable ATP sulfurylase activity but only trace adenosine 5′-phosphosulfate sulfotransferase activity. Both enzyme activities increased in the bundle sheath cells during greening but remained at negligible levels in mesophyll cells. In leaves of maize grown without addition of a sulfur source for 12 d, the specific activity of adenosine 5′-phosphosulfate sulfotransferase and ATP sulfurylase in the bundle sheath cells was higher than in the controls. In the mesophyll cells, however, both enzyme activities remained undetectable. The intercellular distribution of enzymes would indicate that the first two steps of sulfur assimilation are restricted to the bundle sheath cells of C4 plants, and this restriction is independent of ontogeny and the sulfur nutritional status of the plants.  相似文献   

13.
We characterize here the MJ1541 gene product from Methanocaldococcus jannaschii, an enzyme that was annotated as a 5′-methylthioadenosine/S-adenosylhomocysteine deaminase (EC 3.5.4.31/3.5.4.28). The MJ1541 gene product catalyzes the conversion of 5′-deoxyadenosine to 5′-deoxyinosine as its major product but will also deaminate 5′-methylthioadenosine, S-adenosylhomocysteine, and adenosine to a small extent. On the basis of these findings, we are naming this new enzyme 5′-deoxyadenosine deaminase (DadD). The Km for 5′-deoxyadenosine was found to be 14.0 ± 1.2 μM with a kcat/Km of 9.1 × 109 M−1 s−1. Radical S-adenosylmethionine (SAM) enzymes account for nearly 2% of the M. jannaschii genome, where the major SAM derived products is 5′-deoxyadenosine. Since 5′-dA has been demonstrated to be an inhibitor of radical SAM enzymes; a pathway for removing this product must be present. We propose here that DadD is involved in the recycling of 5′-deoxyadenosine, whereupon the 5′-deoxyribose moiety of 5′-deoxyinosine is further metabolized to deoxyhexoses used for the biosynthesis of aromatic amino acids in methanogens.  相似文献   

14.
Escherichia coli DNA ligase (EcoLigA) repairs 3′-OH/5′-PO4 nicks in duplex DNA via reaction of LigA with NAD+ to form a covalent LigA-(lysyl-Nζ)–AMP intermediate (step 1); transfer of AMP to the nick 5′-PO4 to form an AppDNA intermediate (step 2); and attack of the nick 3′-OH on AppDNA to form a 3′-5′ phosphodiester (step 3). A distinctive feature of EcoLigA is its stimulation by ammonium ion. Here we used rapid mix-quench methods to analyze the kinetic mechanism of single-turnover nick sealing by EcoLigA–AMP. For substrates with correctly base-paired 3′-OH/5′-PO4 nicks, kstep2 was fast (6.8–27 s−1) and similar to kstep3 (8.3–42 s−1). Absent ammonium, kstep2 and kstep3 were 48-fold and 16-fold slower, respectively. EcoLigA was exquisitely sensitive to 3′-OH base mispairs and 3′ N:abasic lesions, which elicited 1000- to >20000-fold decrements in kstep2. The exception was the non-canonical 3′ A:oxoG configuration, which EcoLigA accepted as correctly paired for rapid sealing. These results underscore: (i) how EcoLigA requires proper positioning of the nick 3′ nucleoside for catalysis of 5′ adenylylation; and (ii) EcoLigA''s potential to embed mutations during the repair of oxidative damage. EcoLigA was relatively tolerant of 5′-phosphate base mispairs and 5′ N:abasic lesions.  相似文献   

15.
Pseudomonas reinekei MT1 is capable of growing on 4- and 5-chlorosalicylate, involving a pathway with trans-dienelactone hydrolase (trans-DLH) as a key enzyme. It acts on 4-chloromuconolactone formed during cycloisomerization of 3-chloromuconate by hydrolyzing it to maleylacetate. The gene encoding this activity was localized, sequenced and expressed in Escherichia coli. Inductively coupled plasma mass spectrometry showed that both the wild-type as well as recombinant enzymes contained 2 moles of zinc but variable amounts of manganese/mol of protein subunit. The inactive metal-free apoenzyme could be reactivated by Zn2+ or Mn2+. Thus, trans-DLH is a Zn2+-dependent hydrolase using halosubstituted muconolactones and trans-dienelactone as substrates, where Mn2+ can substitute for Zn2+. It is the first member of COG1878 and PF04199 for which a direct physiological function has been reported.  相似文献   

16.
Barley (Hordeum vulgare L.) leaves were used to isolate and characterize the chloroplast NAD(P)H dehydrogenase complex. The stroma fraction and the thylakoid fraction solubilized with sodium deoxycholate were analyzed by native polyacrylamide gel electrophoresis, and the enzymes detected with NADH and nitroblue tetrazolium were electroeluted. The enzymes electroeluted from band S from the stroma fraction and from bands T1 (ET1) and T2 from the thylakoid fraction solubilized with sodium deoxycholate had ferredoxin-NADP oxidoreductase (FNR; EC 1.18.1.2) and NAD(P)H-FeCN oxidoreductase (NAD[P]H-FeCNR) activities. Their NADPH-FeCNR activities were inhibited by 2′-monophosphoadenosine-5′-diphosphoribose and by enzyme incubation with p-chloromercuriphenylsulfonic acid (p-CMPS), NADPH, and p-CMPS plus NADPH. They presented Michaelis constant NADPH values that were similar to those of FNRs from several sources. Their NADH-FeCNR activities, however, were not inhibited by 2′-monophosphoadenosine-5′-diphosphoribose but were weakly inhibited by enzyme incubation with NADH, p-CMPS, and p-CMPS plus NADH. We found that only ET1 contained two polypeptides of 29 and 35 kD, which reacted with the antibodies raised against the mitochondrial complex I TYKY subunit and the chloroplast ndhA gene product, respectively. However, all three enzymes contained two polypeptides of 35 and 53 kD, which reacted with the antibodies raised against barley FNR and the NADH-binding 51-kD polypeptide of the mitochondrial complex I, respectively. The results suggest that ET1 is the FNR-containing thylakoidal NAD(P)H dehydrogenase complex.  相似文献   

17.
Acyl hydrolase activities have been purified from the leaves of Phaseolus multiflorus. The purification procedure involved heat treatment, DEAE-cellulose chromatography, Sephadex G-100 filtration and hexyl agarose chromatography. The elution pattern from hexyl agarose columns together with substrate competition experiments indicated the presence of two hydrolase enzymes. The first could hydrolyse oleoylglycerol and phosphatidylcholine while the second would deacylate glycosylglycerides and oleoylglycerol. Overall purification of both enzymes was ca 70-fold and the MW of the glycosylglyceride-hydrolysing enzyme was in the range 70–78000.  相似文献   

18.
The appearance of enzymes involved in the formation of ureides, allantoin, and allantoic acid, from inosine 5′-monophosphate was analyzed in developing root nodules of soybean (Glycine max). Concomitant with development of effective nodules, a substantial increase in specific activities of the enzymes 5′-nucleotidase (35-fold), purine nucleosidase (10-fold), xanthine dehydrogenase (25-fold), and uricase (200-fold), over root levels was observed. The specific activity of allantoinase remained constant during nodule development. With ineffective nodules the activities were generally lower than in effective nodules; however, the activities of 5′-nucleotidase and allantoinase were 2-fold higher in ineffective nodules unable to synthesize leghemoglobin than in effective nodules. Since the expression of uricase has been shown to be regulated by oxygen (K Larsen, BU Jochimsen 1986 EMBO J 5: 15-19), the expression of the remaining enzymes in the purine catabolic pathway were tested in response to variations in O2 concentration in sterile soybean callus tissue. Purine nucleosidase responded to this treatment, exhibiting a 4-fold increase in activity around 2% O2. 5′-Nucleotidase, xanthine dehydrogenase, and allantoinase remained unaffected by variations in the O2 concentration. Hence, the expression of two enzymes involved in ureide formation, purine nucleosidase and uricase, has been demonstrated to be influenced by O2 concentration.  相似文献   

19.
Recent studies showed that Rai1 and its homologs are a crucial component of the mRNA 5′-end capping quality control mechanism. They can possess RNA 5′-end pyrophosphohydrolase (PPH), decapping, and 5′-3′ exonuclease (toward 5′ monophosphate RNA) activities, which help to degrade mRNAs with incomplete 5′-end capping. A single active site in the enzyme supports these apparently distinct activities. However, each Rai1 protein studied so far has a unique set of activities, and the molecular basis for these differences are not known. Here, we have characterized the highly diverse activity profiles of Rai1 homologs from a collection of fungal organisms and identified a new activity for these enzymes, 5′-end triphosphonucleotide hydrolase (TPH) instead of PPH activity. Crystal structures of two of these enzymes bound to RNA oligonucleotides reveal differences in the RNA binding modes. Structure-based mutations of these enzymes, changing residues that contact the RNA but are poorly conserved, have substantial effects on their activity, providing a framework to begin to understand the molecular basis for the different activity profiles.  相似文献   

20.
The C4 enzyme pyruvate orthophosphate dikinase is encoded by a single gene, Pdk, in the C4 plant Flaveria trinervia. This gene also encodes enzyme isoforms located in the chloroplast and in the cytosol that do not have a function in C4 photosynthesis. Our goal is to identify cis-acting DNA sequences that regulate the expression of the gene that is active in the C4 cycle. We fused 1.5 kb of a 5′ flanking region from the Pdk gene, including the entire 5′ untranslated region, to the uidA reporter gene and stably transformed the closely related C4 species Flaveria bidentis. β-Glucuronidase (GUS) activity was detected at high levels in leaf mesophyll cells. GUS activity was detected at lower levels in bundle-sheath cells and stems and at very low levels in roots. This lower-level GUS expression was similar to the distribution of mRNA encoding the nonphotosynthetic form of the enzyme. We conclude that cis-acting DNA sequences controlling the expression of the C4 form in mesophyll cells and the chloroplast form in other cells and organs are co-located within the same 5′ region of the Pdk gene.  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