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.
Myxobacteria, especially members of the genus Sorangium, are known for their biotechnological potential as producers of pharmaceutically valuable secondary metabolites. The biosynthesis of several of those myxobacterial compounds includes cytochrome P450 activity. Although class I cytochrome P450 enzymes occur wide-spread in bacteria and rely on ferredoxins and ferredoxin reductases as essential electron mediators, the study of these proteins is often neglected. Therefore, we decided to search in the Sorangium cellulosum So ce56 genome for putative interaction partners of cytochromes P450. In this work we report the investigation of eight myxobacterial ferredoxins and two ferredoxin reductases with respect to their activity in cytochrome P450 systems. Intriguingly, we found not only one, but two ferredoxins whose ability to sustain an endogenous So ce56 cytochrome P450 was demonstrated by CYP260A1-dependent conversion of nootkatone. Moreover, we could demonstrate that the two ferredoxins were able to receive electrons from both ferredoxin reductases. These findings indicate that S. cellulosum can alternate between different electron transport pathways to sustain cytochrome P450 activity.The cytochrome P450 (CYP)2 enzymes constitute a superfamily of external monooxygenases. The catalytic versatility of the family members explains their involvement in such diverse biological processes as biosynthesis of steroid hormones, carbon source assimilation, and metabolism of xenobiotics. In addition, cytochrome P450 enzymes have been reported to be involved in the biosynthesis of many pharmaceutically interesting secondary metabolites from a variety of microorganisms (14). Cytochromes P450 are usually dependent on an external electron donor. With respect to their electron transport system they can be divided into several classes, with class I (the mitochondrial/bacterial cytochrome P450 systems) being the predominant form in prokaryotes (5). In this system the electrons required for the enzymatic reaction originate from NAD(P)H and are delivered to the cytochrome P450 via a ferredoxin reductase and a ferredoxin. In a number of examples, the heterologous reconstitution of the electron transfer chain has been shown to be ineffective, if possible at all (5). Thus, it is desirable to identify the natural redox partners, especially if genomic sequence information is available. However, even then the identification of the correct interaction partners remains challenging because the encoding genes are frequently located at genomic loci distant to the cytochrome P450 genes (6, 7). Interestingly, members of both the [2Fe-2S] and the non-[2Fe-2S] ferredoxins have been reported to sustain cytochrome P450 catalyzed reactions. The latter group is further subdivided into mono- and dicluster ferredoxins (i.e. the [3Fe-4S] or [4Fe-4S] and the [3Fe-4S] + [4Fe-4S] or [4Fe-4S] + [4Fe-4S] ferredoxins). Remarkably, cytochrome P450 systems depending on non-[2Fe-2S] ferredoxins have been found exclusively in bacteria to date (8, 9).To fulfill the role as electron mediator, the ferredoxin component of any given cytochrome P450 system has to be reduced. This reduction is achieved by a ferredoxin reductase, which in turn takes up electrons from NAD(P)H. The ferredoxin reductase is often the least characterized constituent of the cytochrome P450 system because these flavoproteins may be unstable (i.e. easily lose their cofactor) and usually show a relatively low level of expression (10).Sorangium cellulosum So ce56 is a genome-sequenced myxobacterial model strain. Because of their biotechnological potential as producers of secondary metabolites, the myxobacteria attract attention from both the academic community and the pharmaceutical industry. To date, more than 100 new basic structures and some 500 derivatives have been reported (11), with almost half of the newly discovered natural products being isolated from the genus Sorangium (11, 12). The potent anti-cancer agent epothilone, for example, was discovered from S. cellulosum So ce90 (13, 14). Epothilone is one of so far seven known myxobacterial compounds, the biosynthesis of which involves cytochromes P450 (15). Besides the epothilones, these are the antifungal leupyrrins (16) and the cytotoxic spirangienes (17) (also from S. cellulosum), the antibiotic myxovirescin from Myxococcus (18), the electron transport inhibitor stigmatellin (19) and the antibiotic aurafuron (20) from Stigmatella aurantiaca, and the antifungal ajudazols from Chondromyces crocatus (21).The recently genome-sequenced myxobacterium S. cellulosum So ce56 (12) shows great potential for biotechnological applications, as judged on the basis of its capacity for the production of secondary metabolites. Three biologically active compounds have been described so far, namely the fungicidal chivosazoles, the macrolide antibiotic etnangien, and the iron chelator myxochelin (12). Moreover, the bioinformatic analysis of the So ce56 genome has revealed numerous biosynthetic gene clusters of yet unknown function (11, 12). With a size of more than 13 Mbp, the genome of S. cellulosum So ce56 is to date the largest sequenced prokaryotic genome (12). It has been shown to harbor 21 cytochrome P450 genes. In light of the significance of S. cellulosum as a viable source of bioactive secondary metabolites (14) and the role of cytochromes P450 in the synthesis of natural products (2), it is of great interest to elucidate the function of these enzymes.Therefore, the investigation of the S. cellulosum So ce56 cytochrome P450 systems opens a fascinating field not only with regard to basic research but also to exploit the biotechnological potential of this model strain. To achieve this goal it is important to provide a functional electron transport chain. Thus, the main objective of this work was to identify a myxobacterial ferredoxin/ferredoxin reductase couple able to support reactions catalyzed by S. cellulosum So ce56 cytochromes P450.  相似文献   

2.
Streptomyces sp. strain 2238-SVT4 produces hatomarubigins A, B, C, and D, which belong to the angucycline family. Among them, hatomarubigin D has a unique dimeric structure with a methylene linkage. PCR using aromatase and cyclase gene-specific primers identified the hrb gene cluster for angucycline biosynthesis in Streptomyces sp. 2238-SVT4. The cluster consisted of 30 open reading frames, including those for the minimal polyketide synthase, ketoreductase, aromatase, cyclase, O-methyltransferase, oxidoreductase, and oxygenase genes. Expression of a part of the gene cluster containing hrbR1 to hrbX in Streptomyces lividans TK23 resulted in the production of hatomarubigins A, B, and C. Hatomarubigin D was obtained from the conversion of hatomarubigin C by a purified enzyme encoded by hrbY, among the remaining genes.The angucycline antibiotics are a large group of naturally occurring aromatic polyketides of microbial origin (11, 15). They exhibit a wide range of biological activities, which include antibacterial, antiviral, antitumor, enzyme inhibitory, and platelet aggregation inhibitory effects. Although all the members contain a benz[a]anthraquinone skeleton of decaketide origin, their structural diversity is very broad and they have a wide variety of oxidation states. Hatomarubigins A, B, C, and D (Fig. (Fig.1)1) belong to the angucycline family and reverse colchicine resistance in multidrug-resistant tumor cells (8). Among them, hatomarubigin D is a unique hatomarubigin C dimer with a methylene linkage. Such a dimer has not been reported previously, and little is known about the mechanism of the methylene bridge formation between two aromatic rings. In this study, a gene cluster for hatomarubigin biosynthesis was identified in Streptomyces sp. strain 2238-SVT4, and a part of the gene cluster was expressed in Streptomyces lividans to produce the hatomarubigins.Open in a separate windowFIG. 1.Structures of angucycline antibiotics.  相似文献   

3.
Cry11Aa and Cyt1Aa of Bacillus thuringiensis are active against mosquitoes and show synergism. Cyt1Aa functions as a membrane receptor inducing Cry11Aa oligomerization. Here we characterized Cry11Aa helix α-3 mutants impaired in oligomerization and toxicity against Aedes aegypti, indicating that oligomerization of Cry11Aa is important for toxin action. Cyt1Aa did not recover the insecticidal activity of Cry11Aa mutants.Bacillus thuringiensis subsp. israelensis has been used worldwide for the control of different mosquitoes that are vectors of several human diseases (10, 11). This bacterium produces different toxins that individually show activity against mosquitoes, i.e., Cry4Aa, Cry4Ba, Cry11Aa, and Cyt1Aa (2). The toxicity of Cry11Aa and Cry4 toxins against Aedes aegypti is greatly increased in the presence of sublethal concentrations of Cyt1Aa (14). Also, Cyt1Aa overcomes the resistance of the Culex quinquefasciatus population to Cry11Aa (12, 13). Cyt1Aa synergizes the toxic activity of Cry11Aa by functioning as a Cry11Aa receptor, facilitating the oligomerization of Cry11Aa and its pore formation activity (7, 8). Oligomerization is a complex event that involves interaction with a toxin receptor and further proteolysis of helix α-1 (3). In the case of the Cry1Ab toxin, helix α-3 of domain I contains coiled-coil structures that are important for oligomerization (4). Some point mutations in helix α-3 do not affect interaction with receptors but severely affected oligomerization, influencing pore formation and toxicity against Manduca sexta larvae (4).Since binding with Cyt1Aa facilitates Cry11Aa oligomerization, we hypothesize that Cry11Aa mutants unable to oligomerize would be affected in synergism with Cyt1Aa and in toxicity. In this report, we analyzed the effect of point mutations in helix α-3 of Cry11Aa on oligomerization, synergism with Cyt1Aa, and toxicity against A. aegypti larvae.Helix α-3 of Cry11Aa potentially forms coiled-coil structures, as determined by the program COILS, which calculates the probability that a sequence will adopt a coiled-coil conformation (6). The coiled-coil structures are characterized by heptads of residues (abcdefg), where positions a and d are occupied mostly by apolar residues and g and e by charged residues. Here we mutagenized some residues located at positions g and a of the predicted coiled-coil (Fig. (Fig.1).1). Substitutions R90E, E97A, Y98E, V104E, and S105E were produced by site-directed mutagenesis (Quick Change; Stratagene, La Jolla, CA) using the pCG6 plasmid (1) as a template and appropriate mutagenic oligonucleotides. Point mutations were verified by automated DNA sequencing at Instituto de Biotecnología-UNAM and transformed into the acrystalliferous B. thuringiensis 407 strain. B. thuringiensis strains were grown in solid nutrient broth sporulation medium supplemented with 10 μg/ml erythromycin (5). Crystal inclusions were purified as described previously (8) and solubilized in 100 mM NaOH for 1 h at 4°C. After solubilization, the Cry11Aa protoxins were dialyzed for 12 h against 50 mM Na2CO3, pH 10.5. The pH was equilibrated at pH 8.6 with equal volumes of 1 M Tris-HCl, pH 8, and protoxins were activated with trypsin (1:50, wt/wt) for 2 h at 25°C. All mutants, with the exception of the V104E mutant, which was not analyzed further, produced crystal inclusions similar to those for the wild-type toxin, composed of a 70-kDa protoxin (Fig. (Fig.2A).2A). After trypsin activation, all mutants produced two polypeptides of 32 and 36 kDa, similarly to the Cry11Aa toxin, suggesting that these mutations did not cause a major structural disturbance (Fig. (Fig.2B).2B). The Cry11Aa and mutant activated toxins were analyzed by circular dichroism spectroscopy (Fig. (Fig.2C).2C). The activated toxins were dialyzed against 10 mM Na2HPO4, 50 mM NaF, pH 9, and then purified by anion-exchange chromatography with HiTrap Q-Sepharose (Pharmacia LKB Biotechnology) in the same buffer, using a linear NaF gradient from 50 to 400 mM. The similarities among the curves indicate that the mutant toxins have a structure similar to that of the wild-type toxin.Open in a separate windowFIG. 1.Schematic representation of the coiled-coil structures of the α-3 helices of Cry1Ab and Cry11Aa toxins. The positions of residues a, b, c, d, e, f, and g of the heptads are presented. The mutated residues in both toxins that affected oligomerization and toxicity are shown in boldface type (reference 4 and this work).Open in a separate windowFIG. 2.SDS-PAGE analysis and circular dichroism spectra of Cry11Aa mutant toxins. (A) The Cry11Aa protoxins were solubilized at pH 10.5 and analyzed by SDS-PAGE (15% acrylamide). (B) SDS-PAGE analysis (15% acrylamide) of the activated toxins with trypsin. Both SDS-polyacrylamide gels were stained with Coomassie blue. Lanes 1, Cry11Aa; lanes 2, E97A mutant; lanes 3, Y98E mutant; lanes 4, R90E mutant; lanes 5, S105E mutant. (C) Analysis of the secondary-structure compositions of the mutants and Cry11Aa activated toxins. Circular dichroism spectra were recorded with a Jasco model J-715 spectropolarimeter equipped with a Peltier temperature control supplied by Jasco. Spectra were collected from 190 to 250 nm. Eight replicate spectra were collected for each sample to improve the signal-to-noise ratios. The final purified-protein concentration was 0.3 mg/ml, and spectra were collected in a 0.1-cm-pathlength cell. The secondary-structure prediction was performed using the CDSSTR algorithm (1a, 11a). Solid black line, Cry11Aa; dotted black line, E97A mutant; dashed black line, Y98E mutant; solid gray line, R90E mutant; dotted gray line, S105E mutant; MRE, mean residue ellipticity; [θ], ellipticity.The toxicity of spore/crystal suspensions of Cry11Aa or the individual mutants (75 to 10,000 ng/ml) was analyzed with bioassays against 10 fourth-instar A. aegypti larvae reared at 28°C, 87% humidity, and 12:12 light-dark conditions in 100 ml dechlorinated water, and mortality was scored after 24 h (four independent assays). The Cry11Aa toxin showed a mean lethal concentration of 355 ng/ml, with 95% confidence limits of 265 to 446 (Probit analysis using Polo-PC LeOra Software). In contrast, the R90E, E97A, Y98E, and S105E mutants were severely affected in toxicity against A. aegypti larvae, since no mortality was observed at the highest concentration used (10,000 ng/ml).We then analyzed the oligomerization of Cry11Aa toxins as previously described (8). Small unilamelar vesicles (SUV), composed of egg yolk phosphatidyl choline, cholesterol (Avanti Polar Lipids, Alabaster, AL), and stearylamine (Sigma, St. Louis, MO) at a 10:3:1 proportion, respectively, were used (8). Cyt1Aa was purified from the 4Q7/pWF45 strain (14) grown as described above. Cyt1Aa inclusions were purified by sucrose gradients, solubilized in 50 mM Na2CO3, 10 mM dithiothreitol, pH 10.5 (2 h at 30°C), and activated with 1:100 proteinase K (Sigma-Aldrich Co.), wt/wt, for 20 min at 30°C.For oligomerization assays, 2.5 μg soluble Cry11Aa or mutant protoxin was incubated for 2 h at 37°C in a 100-μl final volume of 50 mM Na2CO3, pH 10.5, with 200 μM SUV, 1:50 trypsin (wt/wt), and 0.5 μg Cyt1Aa activated toxin. After 2 h of incubation, 1 mM phenylmethylsulfonyl fluoride was added to stop the reaction, and the membrane fraction was separated by centrifugation (1 h at 100,000 × g). The pellet was suspended in the same buffer solution. Oligomeric structures of Cry toxins are highly stable after boiling as well as after urea denaturation (9). The suspension was boiled for 4 min, analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) (8% acrylamide), and electrotransferred onto polyvinylidene difluoride membranes (Millipore, Bedford, MA). The oligomeric and monomeric structures of Cry11Aa were detected using polyclonal anti-Cry11Aa antibody (1/15,000; 1 h) and a secondary antibody coupled with horseradish peroxidase (Sigma, St. Louis, MO) (1/5,000; 1 h) followed by luminol (ECL; Amersham Pharmacia Biotech) as described by the manufacturers. Figure Figure33 shows that only the Cry11Aa wild-type toxin was able to oligomerize, while the mutants were severely impaired in oligomerization.Open in a separate windowFIG. 3.Analysis of Cry11Aa oligomer formation. Soluble Cry11Aa protoxin was activated with trypsin for 2 h at 37°C in the presence of SUV and Cyt1Aa activated toxin. The membrane fraction was separated by ultracentrifugation, and the Cry11Aa protein was analyzed by Western blotting of the membrane pellet with polyclonal anti-Cry11A antibody. The sizes of the proteins were estimated from a molecular prestained plus standard, all blue (Bio-Rad). Lane 1, Cry11Aa; lane 2, R90E mutant; lane 3, Y98E mutant; lane 4, E97A mutant; lane 5, S105E mutant.Finally, the synergistic activity between Cyt1Aa and Cry11Aa was analyzed. A concentration of Cyt1Aa that produced 10% mortality was assayed in the presence of a protein concentration of wild-type Cry11A that produced 20% mortality. Larvae were examined 24 h after treatment, in three repetitions. This particular protein mixture produced a synergism factor of 8. Under these conditions, mortality was more than 80%, due to the synergistic activities of both toxins. Similar experiments were performed with the mutant toxins, using the same concentration of Cyt1Aa toxin and different concentrations (up to 6,000 ng/ml) of the mutant toxins. Cyt1A did not increase the toxicity of the Cry11Aa mutants, since only 10% mortality was observed, even at the highest concentration of the mutant toxins.Previously, helix α-3 of a lepidopteran-specific toxin (Cry1Ab) was subjected to mutagenesis. The R99E and Y107E mutants of the Cry1Ab toxin were severely impaired in oligomerization and toxicity, showing that oligomer formation is a necessary step to kill the larvae (4). The data presented here indicate that oligomer formation is also an essential step in the mechanism of toxicity of the mosquitocidal Cry11Aa toxin and that helix α-3 is involved in this process.  相似文献   

4.
Mutant libraries of avermectin-producing Streptomyces avermitilis strains were constructed by different mutagenesis strategies. A metric was applied to assess the mutation spectrum by calculating the distribution of average phenotypic distance of each population. The results showed for the first time that a microgravity environment could introduce larger phenotype distribution and diversity than UV and N-methyl-N-nitro-N-nitrosoguanidine (NTG) could.Induced mutagenesis is a classical and successful method for improving strains to increase the productivity of commercially significant microbial metabolites. To evaluate different induced-mutagenesis approaches, Klein-Marcuschamer and Stephanopoulos presented a metric based on the quantification of phenotypic diversity to evaluate strain improvement approaches (14).New approaches of inducing mutagenesis emerged with the development of biotechnology, and of these new approaches, spaceflight-induced mutagenesis has led to great progress in strain improvement (6, 15, 26). In outer space, cosmic rays, high vacuum, intense magnetic field, and microgravity induced chromosomal aberrations, which lead to genetic mutations in microorganisms (13). However, it is difficult to carry out spaceflight-induced mutagenesis extensively owing to the limitations of high cost and few chances to board spaceships. Therefore, ground-based simulated experiments have greater practical significance, and high-magnetogravity experiments are a good choice to simulate the space environment (16).Avermectins and its analogues, produced by Streptomyces avermitilis, are major commercial antiparasitic agents for animal health, agriculture, and human infections (7). A variety of mutagenesis methods have been developed to increase the productivity of S. avermitilis (18, 19, 21-23, 25). Though most of them can produce higher mutation rates, the potential of their success in strain improvement is different.In this study, mutant libraries of S. avermitilis strains were constructed by three mutagenesis-inducing strategies: UV, N-methyl-N-nitro-N-nitrosoguanidine (NTG), and high-magnetogravitational environment (HMGE). For each population, the distribution of average phenotypic distance was calculated on the basis of the modified version (15) of the metric of Klein-Marcuschamer and Stephanopoulos (14). The mutation rate was also calculated. A good correlation between the distribution of average phenotypic distance and the percent improvement was found and analyzed. In this way, the potential to produce mutations among different induced-mutagenesis approaches was evaluated to find the most effective one for S. avermitilis.The industrial avermectin-producing S. avermitilis 3-115 strain and the mutants derived from strain 3-115 were grown on YMG agar medium (10). For diversity quantification and preliminary screening, fermentation was carried out in high-throughput format at 28°C. For confirmation of results and secondary screening, mutants that exhibited a higher yield than the wild-type strain were inoculated into shake flasks (10, 11). A high-magnetogravitational experimental platform using the large gradient superconducting magnet was described in detail by Qian et al. (20).The mutant libraries were prepared from S. avermitilis 3-115 by three different mutagenesis strategies. Spore suspensions were prepared in sterile water (106 spores/ml). For UV-induced mutagenesis, 4-ml aliquots of the spore suspension were transferred into sterile petri dishes with a diameter of 80 mm. The petri dishes were then exposed to UV light in a UV-dispensing cabinet fitted with 15-W lamps with about 90% of its radiation at 265 nm. The dishes were placed at a distance of 30.0 cm away from the center of the UV light source and exposed to UV light for 15, 30, 45, 60, 75, and 90 s. The UV-exposed aliquots were then stored in the dark overnight to avoid photoreactivation. For NTG-induced mutagenesis, 9 ml of the spore suspension was added to 1 ml of a sterile solution of NTG (3 mg/ml NTG in phosphate buffer; solution freshly prepared 1 h before use). The samples were shaken at 28°C for 30 min and immediately centrifuged for 10 min at 5,000 rpm, and the supernatant was decanted. The cells were washed three times with sterile water and resuspended in 10 ml of sterile phosphate buffer (pH 7.2). All of the experimental samples were serially diluted with sterile water and plated on YMG plates. For HMGE-induced mutagenesis, the superconducting magnet generated three different magnetic force fields in different places, which corresponded to three apparent levels of gravity (0, 1, and 2 g) and two magnetic induction intensities (12 and 16 T). Specifically, there were three treatment groups in this study: 0-g group (0 g, 12 T), 2-g group (2 g, 12 T), and 1-g group (1 g, 16 T). The YMG plates (plated with 4 ml of spore suspension) were placed at the corresponding places in the platform for 7 days at 28°C to simulate strains grown in space. For all of above induced-mutagenesis experiments, when single colonies were visible on YMG agar medium, they were transferred to 96-well plates for cultivation.To quantify avermectin production, UV absorbance of culture was measured on a multiplate reader (11), and all experiments were repeated twice except where specifically noted. More than 165 clones from each library were screened. The phenotypic distributions of five different populations (including the control) were quantified. We used the optical density at 245 nm (OD245) of the culture as the phenotype for diversity quantification of each mutant library. All data were analyzed with MATLAB (MathWorks, USA). Average phenotypic distance (d) was calculated as follows: (1) where the brackets indicate an average over all pairs of members of the population (colonies i and j) and Pi is the phenotype of colony i and Pj is the phenotype of colony j. In this case, the logarithm of the OD245 was used as the phenotypic measure (Oi) because it was found to be lognormally distributed: (2) The strains from improved wells with respect to the control OD245 were cultured in shake flasks to verify the results. The production of avermectins was determined by high-performance liquid chromatography (HPLC) (Agilent 1200) (8, 11). The positive mutants were defined as the strains that showed increased avermectin B1a production (increased by more than 10%) compared to the original strains. The negative mutants were defined as the strains that showed decreased avermectin B1a production (decreased by more than 10%) compared with the original strains. The mutation rate was calculated as the number of either positive or negative mutants divided by the total number of screened mutants, and the calculation was based on the results of preliminary screening.The distributions of the average phenotypic distances of five different populations were calculated. Bootstrapping was used to derive these distributions, and the results were displayed in a histogram (Fig. (Fig.11 a). To see the statistical significance of the difference between the different groups, the mean and standard deviation of each histogram in Fig. Fig.1a1a were calculated (Fig. (Fig.1b).1b). Homogeneous populations had small average phenotypic distances, whereas diverse populations had larger ones. Larger distance implied larger phenotypic dissimilarity among members of a population. Figure 1a and b showed that the phenotypic diversities in decreasing order were HMGE, NTG, and UV. Among HMGE-induced-mutagenesis libraries, phenotypic diversities in decreasing order were 0-g group (0 g, 12 T), 2-g group (2 g, 12 T), and 1-g group (1 g, 16 T) (Fig. 1a and b). For comparison, the traditional evaluation index, the mutation rate, was also calculated (Fig. (Fig.1c).1c). By using the mutation rate, the positive mutation rate of NTG was the highest, while that of UV was the lowest. Among HMGE-induced-mutagenesis libraries, the mutation rates in decreasing order were 1-g group (1 g, 16 T), 2-g group (2 g, 12 T), and 0-g group (0 g, 12 T) (Fig. (Fig.1c1c).Open in a separate windowFIG. 1.Avermectin production spectrum evaluation of five different mutagenized S. avermitilis populations and the untreated control. Mutations were induced by UV, NTG (N-methyl-N-nitro-N-nitrosoguanidine), or high-magnetogravitational environment (HMGE) treatment. There were three HMGE groups: 0-g group (0 g, 12 T), 2-g group (2 g, 12 T), and 1-g group (1 g, 16 T). (a) Bars in a histogram that reflect the probabilistic distributions of the estimated average phenotypic distance (d in equation 1). (b) To assess the statistical significance of this metric more directly, the mean and standard deviation of each bar in the histogram in panel a were calculated. (c) Mutation rates of the five populations.Figure Figure22 shows the percentages of mutants that exhibited higher yield in both preliminary and secondary screening (percent improved), considering the instability of mutants in induced mutagenesis. The results parallel the findings of the diversity metric (Fig. (Fig.1b),1b), not the mutation rate (Fig. (Fig.1c).1c). To investigate the predictability of the divergence for improved phenotypes, the correlation between the divergence (average phenotype distance) and the occurrence of a mutant with improved production was investigated. Figure Figure33 shows the correlation between divergence and the percent improved. A sigmoid fit and a correlation of R2 = 1 were obtained. The results indicated that for medium divergence (0.6 < divergence < 0.8), improved diversity increased rapidly with the probability of isolating mutants with improved phenotype, while for divergence which was less than 0.6 or more than 0.8, the correlation of divergence with the probability of finding improved mutants was relatively low.Open in a separate windowFIG. 2.Percentage of improved mutants that produce 10% more avermectin B1a than the parent strain. The percentage represents the fraction of “successful” screening events (produce more avermectin B1a on secondary screening) and is a measure of the probability of finding improved mutants in a population.Open in a separate windowFIG. 3.Correlation of divergence and percentage of improved mutants. A sigmoid fit was used (equations shown in the figure).Traditionally, the positive mutation rate is used to evaluate mutation spectrum. However, the positive mutation rate describes only the mutations that occur and cannot describe the extent of mutation or how broad the mutation spectrum is. Therefore, in this study, divergence of mutant libraries was calculated to assess the mutation effect, and divergence was used to evaluate the effect of different induced-mutation strategies, including HMGE, a new spaceflight-simulated mutation strategy. The results of this study indicated the following. (i) HMGE-induced mutagenesis enhanced average phenotype distance and diversity better than UV and NTG mutagenesis did for S. avermitilis. (ii) Microgravity introduced the largest diversity in the genome of S. avermitilis under HMGE conditions. (iii) For medium divergence (0.6 < divergence < 0.8), improved diversity increased the probability of isolating mutants with improved phenotype.NTG was far more efficient than UV irradiation, a conclusion that many other researchers have reached; however, in a study on the efficiency of mutagenesis on spectinomycin resistance in Streptomyces fradiae, it was reported that NTG was more efficient than UV (1-5, 17). The platform of HMGE was developed to simulate the space environment, which has been reported to improve production of certain antibiotics in microorganisms (20). The results indicated that the simulated weightless environment significantly affected cell population and suggested that microgravity may cause the main mutagenic effects on the strains. Results reported here add empirical support to the hypothesis that microgravity is the most important mutagenic factor in spaceflight (12). Microgravity may increase the growth rate of microorganisms, because under microgravity conditions, oxygen in air can be supplied to microorganisms on all surfaces equally, an advantage in the production of biological matter in space (13). It has been hypothesized that microgravity may disturb the system of DNA repair, which blocks or delays the repair of DNA strand breakage. However, some researchers (24) have demonstrated that is not true, and the mutation mechanism is still not clear.Diversity has been reported to be correlated with the probability of finding improved mutants, and improved diversity would increase the probability of isolating mutants with improved phenotype (14). Results from our study partly supported this view. For medium divergence (0.6 < divergence < 0.8), there was a significant correlation between diversity and the probability of isolating mutants with improved phenotype. An optimal mutation rate, which functioned as a balance between uniqueness and retention of function, was proved to exist (9). In addition, those findings demonstrated how optimal error-prone PCR mutation rates may be calculated and indicated that “optimal” rates depended on both the protein and the mutagenesis protocol. Our results concurred with the above findings and showed that no significant correlation was detected for divergence which is less than 0.6 or more than 0.8 but that an optimal mutation rate for medium divergence (0.6 < divergence < 0.8) exists. The findings indicated the existence of a balance between mutation rate and improved phenotype, which implies that too high a mutation rate would cause dysfunction in some gene sequences, while certain mutation rates would produce a few mutated gene sequences helpful for improving productivity.  相似文献   

5.
The gene rapL lies within the region of the Streptomyces hygroscopicus chromosome which contains the biosynthetic gene cluster for the immunosuppressant rapamycin. Introduction of a frameshift mutation into rapL by ΦC31 phage-mediated gene replacement gave rise to a mutant which did not produce significant amounts of rapamycin. Growth of this rapL mutant on media containing added l-pipecolate restored wild-type levels of rapamycin production, consistent with a proposal that rapL encodes a specific l-lysine cyclodeaminase important for the production of the l-pipecolate precursor. In the presence of added proline derivatives, rapL mutants synthesized novel rapamycin analogs, indicating a relaxed substrate specificity for the enzyme catalyzing pipecolate incorporation into the macrocycle.Rapamycin is a 31-member macrocyclic polyketide produced by Streptomyces hygroscopicus NRRL 5491 which, like the structurally related compounds FK506 and immunomycin (Fig. (Fig.1),1), has potent immunosuppressive properties (24). Such compounds are potentially valuable in the treatment of autoimmune diseases and in preventing the rejection of transplanted tissues (16). The biosynthesis of rapamycin requires a modular polyketide synthase, which uses a shikimate-derived starter unit (11, 20) and which carries out a total of fourteen successive cycles of polyketide chain elongation that resemble the steps in fatty acid biosynthesis (2, 27). l-Pipecolic acid is then incorporated (21) into the chain, followed by closure of the macrocyclic ring, and both these steps are believed to be catalyzed by a pipecolate-incorporating enzyme (PIE) (18), the product of the rapP gene (8, 15). Further site-specific oxidations and O-methylation steps (15) are then required to produce rapamycin. Open in a separate windowFIG. 1Structures of rapamycin, FK506, and immunomycin.The origin of the pipecolic acid inserted into rapamycin has been previously established (21) to be free l-pipecolic acid derived from l-lysine (although the possible role of d-lysine as a precursor must also be borne in mind) (9). Previous work with other systems has suggested several alternative pathways for pipecolate formation from lysine (22), but the results of the incorporation of labelled lysine into the pipecolate moiety of immunomycin (Fig. (Fig.1)1) clearly indicate loss of the α-nitrogen atom (3). More recently, the sequencing of the rap gene cluster revealed the presence of the rapL gene (Fig. (Fig.2),2), whose deduced gene product bears striking sequence similarity to two isoenzymes of ornithine deaminase from Agrobacterium tumefaciens (25, 26). Ornithine deaminase catalyzes the deaminative cyclization of ornithine to proline, and we have proposed (15) that the rapL gene product catalyzes the analogous conversion of l-lysine to l-pipecolate (Fig. (Fig.3).3). Open in a separate windowFIG. 2A portion of the rapamycin biosynthetic gene cluster which contains ancillary (non-polyketide synthase) genes (15, 27). PKS, polyketide synthase.Open in a separate windowFIG. 3(A) The conversion of l-ornithine to l-proline by ornithine cyclodeaminase (17). (B) Proposed conversion of l-lysine to l-pipecolic acid by the rapL gene product.Here, we report the use of ΦC31 phage-mediated gene replacement (10) to introduce a frameshift mutation into rapL and the ability of the mutant to synthesize rapamycins in the absence or presence of added pipecolate or pipecolate analogs.  相似文献   

6.
Sulfite dehydrogenases (SDHs) catalyze the oxidation and detoxification of sulfite to sulfate, a reaction critical to all forms of life. Sulfite-oxidizing enzymes contain three conserved active site amino acids (Arg-55, His-57, and Tyr-236) that are crucial for catalytic competency. Here we have studied the kinetic and structural effects of two novel and one previously reported substitution (R55M, H57A, Y236F) in these residues on SDH catalysis. Both Arg-55 and His-57 were found to have key roles in substrate binding. An R55M substitution increased Km(sulfite)(app) by 2–3 orders of magnitude, whereas His-57 was required for maintaining a high substrate affinity at low pH when the imidazole ring is fully protonated. This effect may be mediated by interactions of His-57 with Arg-55 that stabilize the position of the Arg-55 side chain or, alternatively, may reflect changes in the protonation state of sulfite. Unlike what is seen for SDHWT and SDHY236F, the catalytic turnover rates of SDHR55M and SDHH57A are relatively insensitive to pH (∼60 and 200 s–1, respectively). On the structural level, striking kinetic effects appeared to correlate with disorder (in SDHH57A and SDHY236F) or absence of Arg-55 (SDHR55M), suggesting that Arg-55 and the hydrogen bonding interactions it engages in are crucial for substrate binding and catalysis. The structure of SDHR55M has sulfate bound at the active site, a fact that coincides with a significant increase in the inhibitory effect of sulfate in SDHR55M. Thus, Arg-55 also appears to be involved in enabling discrimination between the substrate and product in SDH.Sulfite-oxidizing enzymes protect cells against potentially fatal damage to DNA and proteins caused by exposure to sulfite, and consequently they are found in all forms of life (1). In bacteria, sulfite oxidation is often linked to energy-generating processes during chemolithotrophic growth on reduced sulfur compounds (2, 3), whereas both plant and vertebrate sulfite oxidases have been shown to detoxify sulfite arising from the degradation of methionine and cysteine and exposure to sulfur dioxide (4, 5).All known sulfite-oxidizing enzymes belong to the same family of mononuclear molybdenum enzymes. Their active sites contain one molybdopterin unit per molybdenum atom, and these enzymes may also contain heme groups as accessory redox centers (69). Examples of different types of sulfite-oxidizing molybdoenzymes are the homodimeric plant sulfite oxidase, which does not contain a heme group and uses oxygen as its preferred electron acceptor (9), the homodimeric chicken and human liver sulfite oxidases (CSO3 and HSO, respectively) (10), which are also able to use oxygen as an electron acceptor, and the bacterial sulfite dehydrogenase (SDH) isolated from the soil bacterium Starkeya novella (11, 12), which cannot donate electrons directly to oxygen. Each monomer of CSO and HSO contains a heme b center in addition to the molybdenum center, and the redox centers are located within separate, flexibly linked domains of the same protein subunit. In contrast, the bacterial enzyme is a heterodimer where each subunit of the enzyme contains one redox center. The molybdopterin cofactor is located in the larger 40.2-kDa SorA subunit, and the c-type heme is located in the smaller, 8.8-kDa SorB subunit (12). The SDH quaternary structure thus differs clearly from that of the human and chicken sulfite oxidases.Crystal structures are available for plant sulfite oxidase, CSO, and the bacterial SDH (10, 11, 1315) and have revealed molecular details of the sulfite-oxidizing enzymes. In the CSO structure, the mobile heme b domain occupies a position too removed from the molybdenum active site to mediate efficient electron transfer (10), and indeed the kinetics of this enzyme are known to be complicated by domain movements (16). In contrast, the bacterial SDH is a tight complex with strong electrostatic interactions between the subunits, and the close approach of the redox centers (Mo–Fe distance 16.6 Å) allows for rapid electron transfer (11, 17) (Fig. 1, A and B).Open in a separate windowFIGURE 1.Details of the crystal structure of wild type SDH and comparison with CSO. A, ribbon diagram of the SDH heterodimer with the SorA and SorB subunits colored blue and cyan, respectively, and the redox cofactors in space-filling mode with the molybdenum atom colored green and the iron atom colored violet. B, ribbon diagram of a single subunit of CSO with the molybdopterin binding domain in the same orientation as SorA in A. The cytochrome domain of CSO is clearly in a different position with respect to the molybdenum cofactor than is seen for the cytochrome subunit of SDH. C, SDH molybdopterin cofactor demonstrating the geometry of the molybdenum ligands. The thiol ligands donated by the organic component of molybdopterin and the Cys-104 side chain, and the reactive oxygen ligand (Oeq) sit in the equatorial plane with the axial oxygen (Oax) ligand at the apex of a square pyramid. Atoms are colored as follows: molybdenum (green), sulfur (orange), phosphorous (magenta), oxygen (red), nitrogen (blue), and carbon (yellow in the cofactor and white in the protein). D, hydrogen bonding network around the substrate binding site. The molybdopterin and heme cofactors are shown together with active site residues Cys-104, Arg-55, His-57, Tyr-236, and Gln-33. Figs. 1 and and44 were prepared using Pymol (37).Despite the overall structural differences of these proteins, the coordination geometries of the molybdenum active sites of these sulfite-oxidizing enzymes are nearly identical. The oxidized molybdenum center has a square pyramidal conformation, with three sulfur and two oxo ligands (18). Within this molybdenum center, the equatorial oxo ligand is proposed to be catalytically active, whereas the axial oxo ligand is not thought to participate directly in the reaction (Fig. 1C). During catalysis, the equatorial oxo ligand is transformed into a hydroxy/water ligand as a result of the reduction of the molybdenum center (Fig. 2), and it is in this form that it is generally observed in the CSO and SDH crystal structures.Open in a separate windowFIGURE 2.Proposed reaction mechanism for S. novella sulfite dehydrogenase. The reaction is shown in terms of the redox states of the molybdenum and heme centers present in the enzyme. Shown in boldface type and boxed are the stable redox states of the S. novella SDH. Cyt. c, a mitochondrial type cytochrome c550 (e.g. horse heart or S. novella cytochrome c550) that can act as the external electron acceptor.SDH, CSO, and HSO show similarly high affinities for their substrate, sulfite, and several highly conserved residues surround the substrate-binding and molybdenum active site, namely Tyr-236 (all residues given in SDH numbering (11)), Arg-55, and His-57 (Fig. 1D). Both Arg-55 and Tyr-236 form hydrogen bonds to the catalytically active equatorial Mo-oxo group, whereas His-57 is positioned close to both Arg-55 and Tyr-236 (10, 11) (Fig. 1D). In addition, the crystal structure of the bacterial SDH shows that Arg-55 interacts directly with the second SDH redox center by hydrogen bonding to heme propionate-6 (Fig. 1D) (11).As a result of the similarities in catalytic parameters and the structure of the active site, the bacterial SDH is a very good system for studies of enzymatic sulfite oxidation and especially the molecular basis for catalysis. Since this enzyme does not rely on domain movement for catalysis, it has a less complicated reaction mechanism than the vertebrate enzymes, which facilitates the interpretation of kinetic data, and it can be readily crystallized with both redox centers present in an electron transfer competent conformation. We have previously reported data on the structure, kinetics, EPR, and redox properties of a Y236F-substituted SDH (13). In addition to reduced turnover and substrate affinity, this substitution influences the reactivity of the SDH toward oxygen, turning SDHY236F essentially into an (albeit weak) sulfite oxidase. In order to further understand the roles of the conserved amino acids surrounding the molybdenum active site of sulfite-oxidizing enzymes, we have created two novel amino acid substitutions in the Arg-55 and His-57 residues present at the active site and have investigated their effect on catalytic and spectroscopic parameters of the bacterial SDH. We have also solved the crystal structures of the substituted enzymes, which have provided new insights into the conformation and plasticity of the active site of sulfite-oxidizing enzymes and how the conserved active site residues contribute to sulfite oxidation.  相似文献   

7.
We previously reported that the SLEEPY1 (SLY1) homolog, F-box gene SNEEZY/SLEEPY2 (SNE/SLY2), can partly replace SLY1 in gibberellin (GA) hormone signaling through interaction with DELLAs RGA and GAI. To determine whether SNE normally functions in GA signaling, we characterized the phenotypes of two T-DNA alleles, sne-t2 and sne-t3. These mutations result in no apparent vegetative phenotypes, but do result in increased ABA sensitivity in seed germination. Double mutants sly1-t2 sne-t2 and sly1-t2 sne-t3 result in a significant decrease in plant fertility and final plant height compared to sly1-t2. The fact that sne mutations have an additive effect with sly1 suggests that SNE normally functions as a redundant positive regulator of GA signaling.Key words: gibberellin signaling, GA, SLEEPY1, SNEEZY, DELLA, F-box proteinThis paper describes genetic evidence that the SLEEPY1 (SLY1) homolog SNEEZY/SLEEPY2 (SNE/SLY2) functions redundantly with SLY1 to stimulate gibberellin signaling. GA responses such as seed germination, stem elongation and fertility are promoted by proteolysis of DELLA proteins, negative regulators of the GA signaling.1 In the classic GA signaling model, GA binding to the GA receptor GID1 increases GID1 affinity for DELLA protein. This GID1-GA binding to DELLA causes SLY1, the F-box subunit of an SCF E3 ubiquitin ligase complex, to recognize, bind and ubiquitinate DELLA proteins thereby targeting them for destruction by the 26S proteasome. Thus, loss of SLY1 function results in decreased GA responses, causing dwarfism, delayed flowering, infertility and seed dormancy. The sly1 mutants over-accumulate DELLA proteins due to failure to destroy them through the ubiquitin-proteasome pathway.Overexpression of the SLY1 homolog, SNE, partially rescues the germination, dwarfism and infertility of the sly1-10 mutant.24 SNE overexpression in the sly1-10 background is associated with reduced accumulation of DELLA proteins RGA and GAI, but not of DELLA RGL2. Co-immunoprecipitation assays demonstrated that SNE directly binds RGA protein as well as the cullin subunit of the SCF E3 complex. These recently published data suggest that SNE forms a functional SCF E3 ubiquitin ligase complex that negatively regulates a subset of the DELLA proteins regulated by SLY1.4The finding that SNE overexpression rescues sly1-10 phenotypes through down-regulation of DELLA RGA and GAI suggests that SNE is normally a positive regulator of GA signaling. If this is true, then we expect sne mutations to cause phenotypes resulting from reduced GA response including reduced germination, stature and fertility. To examine this hypothesis, three sne T-DNA mutants were identified: sne-t1, sne-t2 and sne-t3. The sne-t1 allele is a SALK line containing a T-DNA insertion 183-bp upstream of the coding region.5 This line showed no apparent phenotype and was not further characterized. The sne-t2 allele is a Sussman T-DNA line4,6 containing a T-DNA insertion immediately before the ATG that is the SNE translational start codon (Fig. 1A). While this insertion does not disrupt the coding region, it likely disrupts SNE protein translation as the T-DNA contains multiple stop codons. The sne-t3 allele contains a T-DNA insertion within the SNE ORF before amino acid 146 of the 157 amino acid predicted protein (FLAG_461E03).7,8 This allele should result in loss of the last 11 SNE amino acids. We know that loss of the last 8 SLY1 amino acids in sly1-10 results in dwarfism, suggesting that loss of the last 11 SNE amino acids may also cause some loss of function in the small F-box protein. When the homozygous sne-t2 and sne-t3 lines were compared to wild-type Ws, no change was observed either in final plant height or fertility measured in seeds/silique (Fig. 1B). An ABA dose-response curve in seed germination detected a small but reproducible increase in ABA sensitivity during seed germination of sne-t2 and sne-t3 (Fig. 1C). The fact that the sne-t2 and sne-t3 mutants, like sly1-2 and sly1-10, show increased ABA sensitivity suggests that SNE and SLY1 may have similar functions in GA signaling during seed germination.2Open in a separate windowFigure 1The phenotypes of sne-t2 and sne-t3 T-DNA mutants. (A) Schematic diagram of the sne-t2 T-DNA insertion at position −1 bp and of sne-t3 at position +435 bp with respect to the translation start site. (B) Final plant height (upper) and fertility (lower) of indicated genotypes. Letters indicate statistically different classes as determine by t-test. Bars represent standard error. (C) sne mutants show increase in ABA sensitivity. Seeds of wild-type Ws, sne-t2 and sne-t3 were after-ripened for 2 weeks then sown on MS-agar containing indicated concentrations of ABA as described by Steber et al.11 Germination was scored based on radical emergence after incubating 3 days at 4°C followed by 14 days at 22°C. (D) Mutations in SNE cause no significant effect on DELLA RGA, GAI and RGL2 protein accumulation. Total protein was extracted from leaves of 12-d-old seedlings (Top) or flower buds (FB, bottom) and detected as described in Ariizumi et al.4One possible explanation for the lack of apparent GA-insensitive phenotypes in sne T-DNA insertion lines, is that SNE function is redundant with SLY1 in GA signaling.9 If so, we would expect sly1 sne double mutants to show stronger GA-insensitive phenotypes than the sly1 single mutation. Double mutants were constructed containing either the sne-t2 or sne-t3 mutation in the sly-t2 null background. The sly1-t2 allele was chosen because sly1-t2, sne-t2 and sne-t3 are all in the Ws ecotype. The sly1-t2 allele contains a T-DNA insertion within the F-box domain resulting in severe GA-insensitive phenotypes including failure to germinate, reduced stature and infertility.10 The sly1-t2 sne-t2 and sly1-t2 sne-t3 double mutants showed a small but significant decrease in final plant height and fertility (seeds/silique) compared to sly1-t2 (Fig. 1B). This increase in phenotype severity was not associated with an apparent increase in DELLA RGA, GAI or RGL2 protein accumulation (Fig. 1D). It could be that DELLA protein levels in sly1-t2 are so high that any slight increase due to sne mutations is undetectable. Our previous study of SNE overexpression lines showed that SNE has the ability to downregulate RGA and GAI protein accumulation. Figure 1 shows that the chromosomal SNE gene contributes to GA signaling presumably through ubiquitination of DELLA protein.Taken together, the fact that sne mutants show only mild GA-insensitive phenotypes and that the natural SNE expression cannot compensate for lack of SLY1, indicate that SLY1 is the main E3 ubiquitin ligase stimulating GA signaling (this study, reviewed in ref. 4). We cannot rule out the possibility that stronger SNE alleles would show either stronger GA response phenotypes or phenotypes that are unrelated to GA signaling. Indeed, there is evidence to suggest that SNE may have unique functions. The sne-t3 (sne-1) allele results in a shortened root phenotype.8 That SNE is expressed in the endodermis and quiescent center of the root whereas SLY1 is expressed in the stele, suggests that SNE may function independently in the root.8 Moreover, SNE overexpression, but not SLY1 overexpression, results in decreased apical dominance and a prone growth habit suggesting that SNE may play a unique role in development.2,4 Our model is that in addition to regulating DELLA proteins RGA and GAI, SNE may also regulate a yet unidentified target involved in apical dominance (Fig. 2). Future research will need to elucidate the role of SNE in Arabidopsis growth and development.Open in a separate windowFigure 2Model for SNE function in Arabidopsis. Both SLY1 and SNE act as positive regulators of GA responses via DELLA protein destruction. SNE may negatively regulate an unknown protein that maintains apical dominance.  相似文献   

8.
9.
Some AGP molecules or their sugar moieties are probably related to the guidance of the pollen tube into the embryo sac, in the final part of its pathway, when arriving at the ovules. The specific labelling of the synergid cells and its filiform apparatus, which are the cells responsible for pollen tube attraction, and also the specific labelling of the micropyle and micropylar nucellus, which constitutes the pollen tube entryway into the embryo sac, are quite indicative of this role. We also discuss the possibility that AGPs in the sperm cells are probably involved in the double fertilization process.Key words: Arabidopsis, arabinogalactan proteins, AGP 6, gametic cells, pollen tube guidanceThe selective labelling obtained by us with monoclonal antibodies directed to the glycosidic parts of AGPs, in Arabidopsis and in other plant species, namely Amaranthus hypochondriacus,1 Actinidia deliciosa2 and Catharanthus roseus, shows that some AGP molecules or their sugar moieties are probably related to the guidance of the pollen tube into the embryo sac, in the final part of its pathway, when arriving at the ovules. The evaluation of the selective labelling obtained with AGP-specific monoclonal antibodies (Mabs) JIM 8, JIM 13, MAC 207 and LM 2, during Arabidopsis pollen development, led us to postulate that some AGPs, in particular those with sugar epitopes identified by JIM 8 and JIM 13, can be classified as molecular markers for generative cell differentiation and development into male gametes.Likewise, we also postulated that the AGP epitopes recognized by Mabs JIM 8 and JIM 13 are also molecular markers for the development of the embryo sac in Arabidopsis thaliana. Moreover, these AGP epitopes are also present along the pollen tube pathway, predominantly in its last stage, the micropyle, which constitutes the region of the ovule in the immediate vicinity of the pollen tube target, the embryo sac.3We have recently shown the expression of AGP genes in Arabidopsis pollen grains and pollen tubes and also the presence of AGPs along Arabidopsis pollen tube cell surface and tip region, as opposed to what had been reported earlier. We have also shown that only a subset of AGP genes is expressed in pollen grain and pollen tubes, with prevalence for Agp6 and Agp11, suggesting a specific and defined role for some AGPs in Arabidopsis sexual reproduction (Pereira et al., 2006).4Therefore we continued by using an Arabidopsis line expressing GFP under the command of the Agp6 gene promoter sequence. These plants were studied under a low-power binocular fluorescence microscope. GFP labelling was only observed in haploid cells, pollen grains (Fig. 1) and pollen tubes (Fig. 2); all other tissues clearly showed no labelling. These observations confirmed the specific expression of Agp6 in pollen grains and pollen tubes. As shown in the Figures 1 and and2,2, the labelling with GFP is present in all pollen tube extension, so probably, AGP 6 is not one of the AGPs identified by JIM 8 and JIM 13, otherwise GFP light emission would localize more specifically in the sperm cells.5 So we think that MAC 207 which labels the entire pollen tube wall (Fig. 3) may indeed be recognizing AGP6, which seems to be expressed in the vegetative cell. In other words, the specific labelling obtained for the generative cell and for the two male gametes, is probably given by AGPs that are present in very low quantities, apparently not the case for AGP 6 or AGP 11.Open in a separate windowFigure 1Low-power binocular fluorescence microscope image of an Arabidopsis flower with the AGP 6 promoter:GFP construct. The labelling is evident in pollen grains that are being released and in others that are already in the stigma papillae.Open in a separate windowFigure 2Low-power binocular fluorescence microscope image of an Arabidopsis ovary with the AGP6 promoter:GFP construct. The ovary was partially opened to show the pollen tubes growing in the septum, and into the ovules. The pollen tubes are also labelled by GFP.Open in a separate windowFigure 3Imunofluorescence image of a pollen tube growing in vitro, and labeled by MAC 207 monoclonal antibody. The labelling is evident all over the pollen tube wall.After targeting an ovule, the pollen tube growth arrests inside a synergid cell and bursts, releasing the two sperm cells. It has recently been shown that sperm cells, for long considered to be passive cargo, are involved in directing the pollen tube to its target. In Arabidopsis, HAP2 is expressed only in the haploid sperm and is required for efficient pollen tube guidance to the ovules.6 The same could be happening with the AGPs identified in the sperm cells by JIM 8 and JIM 13. We are now working on tagging these AGPs and using transgenic plants aiming to answer to such questions.Pollen tube guidance in the ovary has been shown to be in the control of signals produced by the embryo sac. When pollen tubes enter ovules bearing feronia or sirene mutations (the embryo sac is mutated), they do not stop growing and do not burst. In Zea mays a pollen tube attractant was recently identified in the egg apparatus and synergids.7 Chimeric ZmEA1 fused to green fluorescent protein (ZmEA1:GFP) was first visible within the filiform apparatus and later was localized to nucellar cell walls below the micropylar opening of the ovule. This is the same type of labelling that we have shown in Arabidopsis ovules, using Mabs JIM 8 and JIM 13. We are now involved in the identification of the specific AGPs associated with the labellings that we have been showing.  相似文献   

10.
FtsZ, a bacterial homolog of eukaryotic tubulin, assembles into the Z ring required for cytokinesis. In Escherichia coli, FtsZ interacts directly with FtsA and ZipA, which tether the Z ring to the membrane. We used three-dimensional structured illumination microscopy to compare the localization patterns of FtsZ, FtsA, and ZipA at high resolution in Escherichia coli cells. We found that FtsZ localizes in patches within a ring structure, similar to the pattern observed in other species, and discovered that FtsA and ZipA mostly colocalize in similar patches. Finally, we observed similar punctate and short polymeric structures of FtsZ distributed throughout the cell after Z rings were disassembled, either as a consequence of normal cytokinesis or upon induction of an endogenous cell division inhibitor.The assembly of the bacterial tubulin FtsZ has been well studied in vitro, but the fine structure of the cytokinetic Z ring it forms in vivo is not well defined. Super-resolution microscopy methods including photoactivated localization microscopy (PALM) and three-dimensional-structured illumination microscopy (3D-SIM) have recently provided a more detailed view of Z-ring structures. Two-dimensional PALM showed that Z rings in Escherichia coli are likely composed of loosely-bundled dynamic protofilaments (1,2). Three-dimensional PALM studies of Caulobacter crescentus initially showed that Z rings were comprised of loosely bundled protofilaments forming a continuous but dynamic ring (1–3). However, a more recent high-throughput study showed that the Z rings of this bacterium are patchy or discontinuous (4), similar to Z rings of Bacillus subtilis and Staphylococcus aureus using 3D-SIM (5). Strauss et al. (5) also demonstrated that the patches in B. subtilis Z rings are highly dynamic.Assembly of the Z ring is modulated by several proteins that interact directly with FtsZ and enhance assembly or disassembly (6). For example, FtsA and ZipA promote ring assembly in E. coli by tethering it to the cytoplasmic membrane (7,8). SulA is an inhibitor of FtsZ assembly, induced only after DNA damage, which sequesters monomers of FtsZ to prevent its assembly into a Z ring (9). Our initial goals were to visualize Z rings in E. coli using 3D-SIM, and then examine whether any FtsZ polymeric structures remain after SulA induction. We also asked whether FtsA and ZipA localized in patchy patterns similar to those of FtsZ.We used a DeltaVision OMX V4 Blaze microscope (Applied Precision, GE Healthcare, Issaquah, WA) to view the high-resolution localization patterns of FtsZ in E. coli cells producing FtsZ-GFP (Fig. 1). Three-dimensional views were reconstructed using softWoRx software (Applied Precision). To rule out GFP artifacts, we also visualized native FtsZ from a wild-type strain (WM1074) by immunofluorescence (IF).Open in a separate windowFigure 1Localization of FtsZ in E. coli. (A) Cell with a Z ring labeled with FtsZ-GFP. (B) Rotated view of Z ring in panel A. (C) Cell with a Z ring labeled with DyLight 550 (Thermo Fisher Scientific, Waltham, MA). (D) Rotated view of Z ring in panel C. (B1 and D1) Three-dimensional surface intensity plots of Z rings in panels B and D, respectively. (E) A dividing cell producing FtsZ-GFP. The cell outline is shown in the schematic. (Asterisk) Focus of FtsZ localization; (open dashed ovals) filamentous structures of FtsZ. Three-dimensional surface intensity plots were created using the software ImageJ (19). Scale bars, 1 μm.Both FtsZ-GFP (Fig. 1, A, B, and B1) and IF staining for FtsZ (Fig. 1, C, D, and D1) consistently localized to patches around the ring circumference, similar to the B. subtilis and C. crescentus FtsZ patterns (4,5). Analysis of fluorescence intensities (see Fig. S1, A and B, in the Supporting Material) revealed that the majority of Z rings contain one or more gaps in which intensity decreases to background levels (82% for FtsZ-GFP and 69% for IF). Most rings had 3–5 areas of lower intensity, but only a small percentage of these areas had fluorescence below background intensity (34% for FtsZ-GFP and 21% for IF), indicating that the majority of areas with lower intensity contain at least some FtsZ.To elucidate how FtsZ transitions from a disassembled ring to a new ring, we imaged a few dividing daughter cells before they were able to form new Z rings (Fig. 1 E). Previous conventional microscopy had revealed dynamic FtsZ helical structures (10), but the resolution had been insufficient to see further details. Here, FtsZ visualized in dividing cells by 3D-SIM localized throughout as a mixture of patches and randomly-oriented short filaments (asterisk and dashed oval in Fig. 1, respectively). These structures may represent oligomeric precursors of Z ring assembly.To visualize FtsZ after Z-ring disassembly another way, we overproduced SulA, a protein that blocks FtsZ assembly. We examined E. coli cells producing FtsZ-GFP after induction of sulA expression from a pBAD33-sulA plasmid (pWM1736) with 0.2% arabinose. After 30 min of sulA induction, Z rings remained intact in most cells (Fig. 2 A and data not shown). The proportion of cellular FtsZ-GFP in the ring before and after induction of sulA was consistent with previous data (data not shown) (1,11).Open in a separate windowFigure 2Localization of FtsZ after overproduction of SulA. (A) Cell producing FtsZ-GFP after 0.2% arabinose induction of SulA for 30 min. (B) After 45 min. (B1) Magnified cell shown in panel B. (C) Cell producing native FtsZ labeled with AlexaFluor 488 (Life Technologies, Carlsbad, CA) 30 min after induction; (D) 45 min after induction. (D1) Magnified cell shown in panel D. Scale bars, 1 μm. (Asterisk) Focus of FtsZ localization; (open dashed ovals) filamentous structures of FtsZ.Notably, after 45 min of sulA induction, Z rings were gone (Fig. 2, B and B1), replaced by numerous patches and randomly-oriented short filaments (asterisk and dashed ovals in Fig. 2), similar to those observed in a dividing cell. FtsZ normally rapidly recycles from free monomers to ring-bound polymers (11), but a critical concentration of SulA reduces the pool of available FtsZ monomers, resulting in breakdown of the Z ring (9). The observed FtsZ-GFP patches and filaments are likely FtsZ polymers that disassemble before they can organize into a ring.We confirmed this result by overproducing SulA in wild-type cells and detecting FtsZ localization by IF (Fig. 2, C, D, and D1). The overall fluorescence patterns in cells producing FtsZ-GFP versus cells producing only native FtsZ were similar (Fig. 2, B1 and D1), although we observed fewer filaments with IF, perhaps because FtsZ-GFP confers slight resistance to SulA, or because the increased amount of FtsZ in FtsZ-GFP producing cells might titrate the SulA more effectively.Additionally, we wanted to observe the localization patterns of the membrane tethers FtsA and ZipA. Inasmuch as both proteins bind to the same C-terminal conserved tail of FtsZ (12–14), they would be expected to colocalize with the circumferential FtsZ patches in the Z ring. We visualized FtsA using protein fusions to mCherry and GFP (data not shown) as well as IF using a wild-type strain (WM1074) (Fig. 3 A). We found that the patchy ring pattern of FtsA localization was similar to the FtsZ pattern. ZipA also displayed a similar patchy localization in WM1074 by IF (Fig. 3 B).Open in a separate windowFigure 3Localization of FtsA (A) and ZipA (B) by IF using AlexaFluor 488. (C) FtsA-GFP ring. (D) Same cell shown in panel C with ZipA labeled with DyLight 550. (C1 and D1) Three-dimensional surface intensity plots of FtsA ring from panel C or ZipA ring from panel D, respectively. (E) Merged image of FtsA (green) and ZipA (red) from the ring shown in panels C and D. (F) Intensity plot of FtsA (green) and ZipA (red) of ring shown in panel E. The plot represents intensity across a line drawn counterclockwise from the top of the ring around the circumference, then into its lumen. Red/green intensity plot and three-dimensional surface intensity plots were created using the software ImageJ (19). Scale bar, 1 μm.To determine whether FtsA and ZipA colocalized to these patches, we used a strain producing FtsA-GFP (WM4679) for IF staining of ZipA using a red secondary antibody. FtsA-GFP (Fig. 3 C) and ZipA (Fig. 3 D) had similar patterns of fluorescence, although the three-dimensional intensity profiles (Fig. 3, C1 and D1) reveal slight differences in intensity that are also visible in a merged image (Fig. 3 E). Quantitation of fluorescence intensities around the circumference of the rings revealed that FtsA and ZipA colocalized almost completely in approximately half of the rings analyzed (Fig. 3 F, and see Fig. S2 A), whereas in the other rings there were significant differences in localization in one or more areas (see Fig. S2 B). FtsA and ZipA bind to the same C-terminal peptide of FtsZ and may compete for binding. Cooperative self-assembly of FtsA or ZipA might result in large-scale differential localization visible by 3D-SIM.In conclusion, our 3D-SIM analysis shows that the patchy localization of FtsZ is conserved in E. coli and suggests that it may be widespread among bacteria. After disassembly of the Z ring either in dividing cells or by excess levels of the cell division inhibitor SulA, FtsZ persisted as patches and short filamentous structures. This is consistent with a highly dynamic population of FtsZ monomers and oligomers outside the ring, originally observed as mobile helices in E. coli by conventional fluorescence microscopy (10) and by photoactivation single-molecule tracking (15). FtsA and ZipA, which bind to the same segment of FtsZ and tether it to the cytoplasmic membrane, usually display a similar localization pattern to FtsZ and each other, although in addition to the differences we detect by 3D-SIM, there are also likely differences that are beyond its ∼100-nm resolution limit in the X,Y plane.As proposed previously (16), gaps between FtsZ patches may be needed to accommodate a switch from a sparse Z ring to a more condensed ring, which would provide force to drive ring constriction (17). If this model is correct, the gaps should close upon ring constriction, although this may be beyond the resolution of 3D-SIM in constricted rings. Another role for patches could be to force molecular crowding of low-abundance septum synthesis proteins such as FtsI, which depend on FtsZ/FtsA/ZipA for their recruitment, into a few mobile supercomplexes.How are FtsZ polymers organized within the Z-ring patches? Recent polarized fluorescence data suggest that FtsZ polymers are oriented both axially and circumferentially within the Z ring in E. coli (18). The seemingly random orientation of the non-ring FtsZ polymeric structures we observe here supports the idea that there is no strong constraint requiring FtsZ oligomers to follow a circumferential path around the cell cylinder. The patches of FtsZ in the unperturbed E. coli Z ring likely represent randomly oriented clusters of FtsZ filaments that are associated with ZipA, FtsA, and essential septum synthesis proteins. New super-resolution microscopy methods should continue to shed light on the in vivo organization of these protein assemblies.  相似文献   

11.
In plants, the division of peroxisomes is mediated by several classes of proteins, including PEROXIN11 (PEX11), FISSION1 (FIS1) and DYNAMIN-RELATED PROTEIN3 (DRP3). DRP3A and DRP3B are two homologous dynamin-related proteins playing overlapping roles in the division of both peroxisomes and mitochondria, with DRP3A performing a stronger function than DRP3B in peroxisomal fission. Here, we report the identification and characterization of the peroxisome division defective 2 (pdd2) mutant, which was later proven to be another drp3A allele. The pdd2 mutant generates a truncated DRP3A protein and exhibits pale green and retarded growth phenotypes. Intriguingly, this mutant displays much stronger peroxisome division deficiency in root cells than in leaf mesophyll cells. Our data suggest that the partial GTPase effector domain retained in pdd2 may have contributed to the distinct mutant phenotype of this mutant.Key words: peroxisome division, dynamin-related protein, arabidopsisIn eukaryotic cells, peroxisomes are surrounded by single membranes and house a variety of oxidative metabolic pathways such as lipid metabolism, detoxification and plant photorespiration.1,2 To accomplish multiple tasks, the morphology, abundance and positioning of peroxisomes need to be highly regulated. Three families of proteins, whose homologs are present across different kingdoms, have been shown to be involved in peroxisome division in Arabidopsis. The PEX11 protein family is composed of five integral membrane proteins with primary roles in peroxisome elongation/tubulation, the initial step in peroxisome division.35 Although the exact function of PEX11s has not been demonstrated, these proteins are believed to participate in peroxisome membrane modification.6,7 The FIS1 family consists of two isoforms, which are C-terminal tail-anchored membrane proteins with rate limiting functions at the fission step.8,9 DRP3A and DRP3B belong to a superfamily of dynamin-related proteins, which are large and self-assembling GTPases involved in the fission and fusion of membranes by acting as mechanochemical enzymes or signaling GTPases.10 The function of PEX11 seems to be exclusive to peroxisomes, whereas DRP3 and FIS1 are shared by the division machineries of both peroxisomes and mitochondria in Arabidopsis.8,9,1116 FIS1 proteins are believed to tether DRP proteins to the peroxisomal membrane,17,18 but direct evidence has not been obtained from plants. DRP3A and DRP3B share 77% sequence identity at the protein level and are functionally redundant in regulating mitochondrial division; however, DRP3A''s role on the peroxisome seems stronger and cannot be substituted by DRP3B in peroxisome division.8,13,15In a continuous effort to identify components of the plant peroxisome division apparatus from Arabidopsis, we performed genetic screens in a peroxisomal marker background expressing the YFP (yellow fluorescent protein)-PTS1 (peroxisome targeting signal 1, containing Ser-Lys-Leu) fusion protein. Mutants with defects in the morphology and abundance of fluorescently labeled peroxisomes are characterized. Following our analysis of the pdd1 mutant, which turned out to be a strong allele of DRP3A,8 we characterized the pdd2 mutant.In root cells of the pdd2 mutant, extremely elongated peroxisomes and a beads-on-a-string peroxisomal phenotype are frequently observed (Fig. 1A and B). These peroxisome phenotypes resemble those of pdd1 and other strong drp3A alleles previously reported.8,15 However, the peroxisome phenotype seems to be less dramatic in leaf mesophyll cells. For instance, in addition to the decreased number of total peroxisomes, peroxisomes in leaf cells are only slightly elongated or exhibit a beads-on-a-string phenotype (Fig. 1C and D). Previously, we reported the phenotypes of three strong drp3A alleles, all of which contain a large number of peroxules, long and thin membrane extensions from the peroxisome,8 yet such peroxisomal structures are not observed in pdd2. On the other hand, pdd2 has a more severe growth phenotype than most drp3A alleles, as it is slow in growth and has pale green leaves (Fig. 1E). Genetic analysis showed that pdd2 segregates as a single recessive mutation (data not shown).Open in a separate windowFigure 1Phenotypic analyses of pdd2 and identification of the PDD2 gene. (A–D) Confocal micrographs of root and mesophyll cells in 3-week-old wild type and pdd2 mutant plants. Green signals show peroxisomes; red signals show chloroplasts. Scale bars = 20 µm. (E) Growth phenotype of 3-week-old mutants. (F) Map-based cloning of the PDD2 gene. Genetic distance from PDD2 is shown under each molecular marker. Positions for mutations in previously analyzed drp3A alleles and pdd2 are indicated in the gene schematic. drp3A-1 and drp3A-2 are T-DNA insertion mutants, whereas pdd1 is an EMS mutant containing a premature stop codon in exon 6. (G) A schematic of the DRP3A (PDD2) protein with functional domains indicated. The pdd2 allele encodes a truncated protein lacking part of the GED domain.The unique combination of peroxisomal and growth phenotypes of pdd2 prompted us to use map-based cloning to identify the PDD2 gene, with the hope to discover novel proteins in the peroxisome division machinery. A population of approximately 6,000 F2 plants (pdd2 × Ler) was generated. After screening 755 F2 mutants, the pdd2 mutation was mapped to the region between markers T10C21 and F4B14 on the long arm of chromosome 4 (Fig. 1F). Since this region contains DRP3A, we sequenced the entire DRP3A gene in pdd2 and identified a G→A transition at the junction of the 18th exon and intron (Fig. 1F). Further analysis revealed that the point mutation at this junction caused mis-splicing of intron 18, introducing a stop codon in the GTPase effector domain GED near the C terminus (Fig. 1G).DRPs share with the classic dynamins an N-terminal GTPase domain, a middle domain (MD), and a regulatory motif named the GTPase effector domain (GED) (Fig. 1G). To date, a total of 26 drp3A mutant alleles carrying missense or nonsense mutations along the length of the DRP3A gene have been isolated.8,15 The combined peroxisomal and growth phenotype of pdd2 and the nature of the mutation in this allele are unique among all the drp3A alleles, indicating that the partial GED domain retained in pdd2 may have created some novel function for this protein. Further analysis of the truncated protein may be necessary to test this prediction.  相似文献   

12.
Reduced ferredoxin is an intermediate in the methylotrophic and aceticlastic pathway of methanogenesis and donates electrons to membrane-integral proteins, which transfer electrons to the heterodisulfide reductase. A ferredoxin interaction has been observed previously for the Ech hydrogenase. Here we present a detailed analysis of a Methanosarcina mazei Δech mutant which shows decreased ferredoxin-dependent membrane-bound electron transport activity, a lower growth rate, and faster substrate consumption. Evidence is presented that a second protein whose identity is unknown oxidizes reduced ferredoxin, indicating an involvement in methanogenesis from methylated C1 compounds.The aceticlastic pathway of methanogenesis creates approximately 70% (10) of the biologically produced methane and is of great ecological importance, as methane is a potent greenhouse gas. Organisms using this pathway to convert acetate to methane belong exclusively to the genera Methanosarcina and Methanosaeta. The two carbon atoms of acetate have different fates in the pathway. The methyl moiety is converted to methane, whereas the carbonyl moiety is further oxidized to CO2 and the electrons derived from this oxidation step are used to reduce ferredoxin (Fd) (6). During methanogenesis from methylated C1 compounds (methanol and methylamines), one-quarter of the methyl groups are oxidized to obtain electrons for the reduction of heterodisulfide (27). A key enzyme in the oxidative part of methylotrophic methanogenesis is the formylmethanofuran dehydrogenase, which oxidizes the intermediate formylmethanofuran to CO2 (7). The electrons are transferred to Fd. It has been suggested that reduced ferredoxin (Fdred) donates electrons to the respiratory chain with the heterodisulfide (coenzyme M [CoM]-S-S-CoB) as the terminal electron acceptor and that the reaction is catalyzed by the Fdred:CoM-S-S-CoB oxidoreductase system (7, 24). The direct membrane-bound electron acceptor for Fdred is still a matter of debate; for the Ech hydrogenase, a reduced ferredoxin-accepting, H2-evolving activity has been observed for Methanosarcina barkeri (20), which implies that the H2:CoM-S-S-CoB oxidoreductase system is involved in electron transport (13). Direct electron flow from the Ech hydrogenase to the heterodisulfide reductase has not been shown to date (20, 21). In contrast to M. barkeri, Methanosarcina acetivorans lacks the Ech hydrogenase (11). It can nevertheless grow on acetate, which is why another complex present in this organism, the Rnf complex, is thought to be involved in the aceticlastic pathway of methanogenesis as an acceptor for Fdred (8, 10, 17). The Methanosarcina mazei genome, however, contains genes coding for the Ech hydrogenase, but this species lacks the Rnf complex (5).To investigate whether the Ech hydrogenase is the only means by which M. mazei channels electrons from Fdred into the respiratory chain, a mutant lacking the Ech hydrogenase (M. mazei Δech mutant) was constructed. Electron transport experiments using Fdred as the electron donor and CoM-S-S-CoB as the electron acceptor were conducted with wild-type and mutant membranes to gain deeper insight into the actual membrane-bound protein complexes that accept electrons from Fdred. Furthermore, an in-depth characterization of the growth and trimethylamine (TMA) consumption of the Δech mutant was performed, which provided insight into the in vivo role of Ech hydrogenase.  相似文献   

13.
14.
15.
Relaxation of a hERG K+ channel model during molecular-dynamics simulation in a hydrated POPC bilayer was accompanied by transitions of an arginine gating charge across a charge transfer center in two voltage sensor domains. Inspection of the passage of arginine side chains across the charge transfer center suggests that the unique hydration properties of the arginine guanidine cation facilitates charge transfer during voltage sensor responses to changes in membrane potential, and underlies the preference of Arg over Lys as a mobile charge carrier in voltage-sensitive ion channels.The response of voltage-sensitive ion channels to changes in membrane potential is mediated by voltage sensor domains (VSD) containing a transmembrane helical segment (S4) with a repeating motif of positively charged and hydrophobic amino acids (Fig. 1) (1,2). Changes in membrane potential drive the S4 helix through the membrane plane with the charged side chains (largely arginine) on S4 swapping Glu/Asp carboxylate partners that lie on less mobile elements of the VSD (2). Movement of S4 is coupled to the ion-conducting pore to transmit changes in membrane potential to channel gating (3).Open in a separate windowFigure 1Structures of the VSD of membrane domains before MD in a POPC bilayer. The S2 (pink) and S4 (blue) helices of the VSD of the hERG model (A) and Kv1.2/2.1 chimera structure (B) are highlighted. (C) Sequence alignment of S2 and S4 among homologous voltage-sensitive K+ channels.The VSD charge-pairing motif of K+ and Na+ channels is best represented in VSD states at zero membrane potential (S4 helix up) for which crystal structures exist for Kv1.2 (4), Kv1.2/2.1 chimera (5), and Nav channels (6,7). In these states, positively charged residues on the intra- and extracellular sections of the S4 helix are separated by a hydrophobic charge-transfer center (CTC) (1) or plug (8) containing a highly conserved Phe residue (Fig. 1). This plug restricts water incursion across the VSD, focusing the electric field across a narrow region near the bilayer center. In voltage-driven transitions between S4 down- and up-states, positively charged S4 side chains move across the CTC.The ether-à-go-go (eag) and eag-related family of voltage-sensitive K+ channels likely share similar charge pairing interactions with VSDs in other channels (9,10). However, eag VSDs contain an extra negative charge on S2 (underlined in Fig. 1 C) so that in hERG, Asp residues (D460 and D466) lie approximately one helical turn above and below the conserved charge-transfer center Phe (F463) (Fig. 1). This eag-specific motif might be expected to facilitate transfer of Arg side chains through the CTC and to stabilize the voltage sensor (VS) in the up state. We recently described an open state (VS-up) hERG model built on the crystal structure template of the Kv1.2/2.1 chimera and molecular-dynamics (MD) simulation of this model in a hydrated POPC bilayer (11). We have inspected an extended version of this simulation and identified transitions of a gating charge into the CTC despite the absence of a membrane potential change. These transitions are absent in equivalent MD simulations of the chimera structure in a POPC bilayer.Fig. 1 shows a single VS from starting structures of the hERG model and the chimera structure in a hydrated POPC bilayer, after restrained MD to anneal the protein-lipid interface (see Methods in the Supporting Material). Because the hERG model is constructed on the chimera structure according to the alignment in Fig. 1 the pattern of pairing between S4 charges and acidic VS side chains is equivalent in the hERG model and chimera structure.The arrangement of charge-paired side chains remains constant during MD in all subunits of the chimera (e.g., Fig. 2 E and see Fig. S2 in the Supporting Material). However, in two subunits of the hERG model the R534 side chain moves toward the extracellular side of the bilayer, sliding into the CTC to form a charge interaction with the extra Asp residue (D460 in hERG) that lies just above F463 (Fig. 2, AC). This transition is facilitated by changes in side-chain rotamers of R534 and F463 as the planar Arg guanidine group rotates past the F463 ring, and the availability of D460 as a counterion for the R534 guanidine (Fig. 2). Movement of an Arg guanidine past the Phe side chain of the CTC is similar to that described in steered MD of an isolated VS domain (12).Open in a separate windowFigure 2Movement of the R534 side chain across the CTC in chain a of the hERG model simulation (A). Similar transitions are observed in chains a and b (panels B and C), but not chains c (D) or d (not shown), where the R534 side chain remains close to D466. In all subunits of the Kv1.2/2.1 chimera simulation, charge pairing of the starting structure (Fig. 1B) was maintained throughout (e.g., panel E and see Fig. S2 in the Supporting Material). (Black and blue lines) Distances from the Arg CZ or Lys ε atom to the two O atoms, respectively, of Asp or Glu.Mason et al. (13) have shown, using neutron scattering, that the low charge density guanidine cation (Gdm+) corresponding to the Arg side chain is poorly hydrated above and below the molecular plane. This property may underlie the universal preference for Arg (over Lys) in voltage sensor charge transfer. Although the poorly-hydrated surfaces of Gdm+ interact favorably with nonpolar (especially planar) surfaces (14,15), Gdm+ retains in-plane hydrogen bonding (13). In the transition of R534 across the CTC, in-plane solvation of the guanidine side chain is provided initially by D466, D501, and water molecules below the CTC, and during and after the transition by D501 and D460 side chains and waters above the CTC (Fig. 3, A and B). Complete transfer of the R534 side chain across the CTC was not observed, but would be expected to involve movement of the guanidine group away from H-bonding distance with D501.Open in a separate windowFigure 3In-plane solvation of R534 guanidine in the charge transfer center during the hERG model MD (A). (Dotted lines) H-bond distances of <2.5 Å. The right-hand group consists of top-down (B) and end-on (C) views of the distribution of oxygen atoms around the side chain of hERG R534 at 20-ns intervals during MD (subunit a). (D) End-on view of equivalent atom distributions around the K302 side chain during the Kv1.2/2.1 chimera MD (subunit c). (Red spheres, water O; pink, Asp OD1 and OD2; purple:, Glu OE1 and OE2.)The atom distribution around the R534 side chain during MD (Fig. 3, B and C) conforms to the experimental Gdm+ hydration structure (13), with H-bonding to waters and side-chain Asp O atoms exclusively in the guanidine plane. The passage of Gdm+ through the CTC is facilitated by the hydrophobic nature of Gdm+ above and below the molecular plane (13), which allows interaction with the nonpolar groups (especially F463) in the CTC (Fig. 3 A and see Fig. S3). This contrasts with the solvation properties of the Lys amino group (e.g., K302 of the Kv1.2/2.1 chimera (Fig. 1), which has a spherical distribution of H-bonding and charge-neutralizing oxygen atoms (Fig. 3 D and see Fig. S4).To further test these interpretations, we ran additional MD simulations of the isolated hERG VS domain model and an R534K mutant in a hydrated POPC bilayer. Again, the R534 side chain entered the CTC in the wild-type model simulation whereas the K534 side chain did not (see Fig. S5). Inspection of the atom distributions in Fig. 3 D (and see Fig. S4) indicates that the pocket below the conserved Phe of the CTC is particularly favorable for a Lys side chain, with waters and acidic side chains that satisfy the spherical solvation requirements of the terminal amino group, and nonpolar side chains that interact with the aliphatic part of the side chain.The occurrence of transitions of the R534 side chain through the CTC in the hERG model, in the absence of a change in membrane potential, indicates a relaxation from a less-stable starting structure. However, the path of the R534 side chain provides useful molecular-level insight into the nature of charge transfer in voltage sensors. How do these observations accord with broader evidence of charge transfer in voltage-sensitive channels in general, and hERG in particular? Studies with fluorinated analogs of aromatic side chains equivalent to F463 of hERG or F233 of the chimera indicate the absence of a significant role for cation-π interactions involving the CTC aromatic group in K+ and Nav channels, although a planar side chain is preferred in some cases (1,16). In hERG, F463 can be replaced by M, L, or V with small effects on channel gating (17), indicating that the hERG CTC requires only a bulky nonpolar side chain to seal the hydrophobic center of the VS and allow passage of the Arg side chain through the CTC. Both absence of requirement for cation-π interactions, and accommodation of nonplanar hydrophobic side chains in a functional hERG CTC, are broadly consistent with the interpretation that it is the poorly-hydrated nature of the Arg guanidine group above and below the molecular plane (together with its tenacious proton affinity (18)) that governs its role in carrying gating charge in voltage sensors.While the simulations suggest that R534 may interact with D460 in the open channel state, the possibility that the extra carboxylate side chain above the CTC might facilitate gating charge transfer is seemingly inconsistent with the slow activation of hERG, although hERG D460C does activate even more slowly than the WT channel (9). However, S4 movement in hERG occurs in advance of channel opening (19), and slow gating is partly mediated by interactions involving hERG cytoplasmic domains (20); thus, slow S4 movement may not be an inherent property of the hERG voltage sensor. Recent studies show that when hERG gating is studied at very low [Ca2+] (50 μM) and low [H+] (pH 8.0), the channel is strongly sensitized in the direction of the open state; this effect is reduced in hERG D460C (and hERG D509C) (10). These observations support a role for the extra hERG Asp residues in binding Ca2+ (and H+) (10), allowing the channel to be allosterically responsive to changes in pH and [Ca2+]. A true comparison of a hERG model with experimental channel gating might involve studies on a channel lacking cytoplasmic domains that modulate gating, and using conditions (high pH and low [Ca2+]) that leave the eag-specific Asp residues unoccupied. This could reveal the inherent current-voltage relationships and kinetics of the hERG voltage sensor.  相似文献   

16.
Jasmonate (JA) inhibits root growth of Arabidopsis thaliana seedlings. The mutation in COI1, that plays a central role in JA signaling, displays insensitivity to JA inhibition of root growth. To dissect JA signaling pathway, we recently isolated one mutant named psc1, which partially suppresses coi1 insensitivity to JA inhibition of root growth. As we identified the PSC1 gene as an allele of DWF4 that encodes a key enzyme in brassinosteroid (BR) biosynthesis, we hypothesized and demonstrated that BR is involved in JA signaling and negatively regulates JA inhibition of root growth. In our Plant Physiology paper, we analyzed effects of psc1 or exogenous BR on the inhibition of root growth by JA. Here we show that treatment with brassinazole (Brz), a BR biosynthesis inhibitor, increased JA sensitivity in both coi1-2 and wild type, which further confirms that BR negatively regulates JA inhibition of root growth. Since effects of psc1, Brz and exogenous BR on JA inhibition of root growth were mild, we suggests that BR negatively finely regulates JA inhibition of root growth in Arabidopsis.Key words: jasmonate signaling, root growth, brassinosteroid, brassinazole, arabidopsisJasmonate (JA) regulates many plant developmental processes and stress responses.1,2 COI1 plays a central role in JA signaling and is required for all JA responses in Arabidopsis.3,4 coi1-1, a strong mutation in COI1, is male sterile and exhibits loss of all JA responses tested to date, such as JA inhibition of root growth, the expression of JA-induced genes, and susceptibility to insect attack and pathogen infection, and coi1-2, a weak mutant of COI1, shows similar JA responses to coi1-1 except for partially fertile that makes it able to produce a small quantity of seeds.5To investigate COI1-mediated JA responses and dissect JA signaling pathway, we conducted genetic screens for suppressors of coi1-2. Previously, we identified cos1 that completely suppresses coil-2 insensitive to JA.6 Recently, we isolated the psc1 mutant that partially suppresses coi1-2 insensitivity to JA, and found that PSC1 is an allele of DWF4.7Since the DWF4 gene encodes a key enzyme in brassinosteroid (BR) biosynthesis,8 we hypothesized that BR is involved in JA signaling. By physiological analysis, we showed that psc1 partially restored JA inhibition of root growth in coi1-2 background and displayed JA hypersensitivity in wild-type COI1 background, the effects of psc1 were eliminated by exogenous BR, and that exogenous BR could attenuated JA inhibition of root growth in wild type. These findings demonstrated that BR is involved in JA signaling and indicated that BR negatively regulates JA inhibition of root growth.BR is a family of polyhydroxylated steroid hormones involved in many aspects of plant growth and development. The BR-deficient mutants exhibited severely retarded growth that was able to be rescued by exogenous BR.9 Brassinazole (Brz) is a BR biosynthesis inhibitor. The Arabidopsis seedlings treated with Brz displayed a BR deficient-mutant-like phenotype, which could be elimilated by exogenous BR.10To determine wether treatment with Brz affects JA inhibition of root growth, the seedlings of wild type and coi1-2 were grown in MS medium supplemented with MeJA and/or Brz. As shown in Figure 1, the relative root length was obviously reduced in both coi1-2 and wild type when treated with Brz relative to without Brz, indicating that the repression of BR biosynthesis by Brz could increase JA sensitivity. These results further confirm BR negatively regulates JA inhibition of root growth.Open in a separate windowFigure 1Effect of Brz on JA inhibition of root growth. Brz increased JA inhibition of root growth in both coi1-2 and wild type (WT). Root length of 7-day-old seedlings grown in MS medium containing 0, 5 and 10 μM MeJA without (−) or with (+) 0.5 μM Brz was expressed as a percentage of root length in MS without (−) or with (+) 0.5 µM Brz. Error bars represent SE (n > 30).It has been demonstrated that JA connects with other plant hormones including auxin, ethylene, abscisic acid, salicylic acid and gibberellin to form complex regulatory networks modulating plant developmental and stress responses.1115 We found that BR negatively regulates JA inhibition of root growth, suggesting that a cross talk between JA and BR exists in planta, which extends our understandings on the JA signal transduction.COI1 is a JA receptor16 and DWF4 catalyzes the rate-limiting step in BR-biosynthesis pathway.8 We found that JA inhibits DWF4 expression, this inhibition was dependent on COI1,7 indicating that DWF4 is downregulated by JA and is located downstream of COI1 in the JA signaling pathway.Since the effects of psc1, Brz, and exogenous BR on JA inhibition of root growth were mild, and the DWF4 expression was partially repressed by JA (Ren et al. 2009, Fig. 1), we suggest that BR negatively finely regulates JA inhibition of root growth, and propose a model for these regulations. As shown in Figure 2A, JA signal passes COI1 repressing substrates, such as JAZs,17,18 i.e., JA activates degradation of substrates via SCFCOI1-26S proteasome,1618 whereas substrates positively regulate root growth through other regulators. JA also partially inhibits DWF4 expression through COI1, reducing BR that is required for root growth.7,9 Mutation in COI1 interrupts JA signaling for failing in degradation of substrates and repression of DWF4 as well, resulting in JA-insensitivity (Fig. 2B). However, mutation in DWF4 or treatment with Brz causes a reduction in BR, which affects root growth, leading to JA-hypersensitivity in wild-type COI1 background (Fig. 2C and E) and partial restoration of JA sensitivity in coi1-2 background (Fig. 2D and F). Whereas, an application of exogenous BR could eliminate the effect of BR reduction resulted from repression of DWF4 by JA on root growth, attenuating JA sensitivity in wild type (Fig. 2G). Because the inhibition of DWF4 expression by JA is dependent on COI1, the coi1 mutant treated with exogenous BR do not show alteration in JA sensitivity (Fig. 2H).Open in a separate windowFigure 2A model for that BR negatively finely regulates JA inhibition of root growth in Arabidopsis. (A–D) Treatment with JA in wild type (A), coi1-2 (B), psc1 (C) and psc1coi1 (D). (E and F) Treatments with JA and Brz in wild type (E) and coi1-2 (F). (G and H) Treatments with JA and exogenous BR in wild type (G) and coi1-2 (H). Arrows indicate positive regulation or enhancement, whereas blunted lines indicate repression or negative regulation. Crosses indicate interruption or impairment. The letter “S” indicates substrates of SCFCOI1. Thicker arrows and blunted lines represent the central JA signaling pathway regulating JA inhibition of root growth. Broken arrows represent JA signaling pathway in which other regulators are involved. The intensity of gray boxes represents the degree of JA inhibition on root growth.  相似文献   

17.
A novel lachrymatory factor synthase (LFS) was isolated and purified from the roots of the Amazonian medicinal plant Petiveria alliacea. The enzyme is a heterotetrameric glycoprotein comprised of two α-subunits (68.8 kD each), one γ-subunit (22.5 kD), and one δ-subunit (11.9 kD). The two α-subunits are glycosylated and connected by a disulfide bridge. The LFS has an isoelectric point of 5.2. It catalyzes the formation of a sulfine lachrymator, (Z)-phenylmethanethial S-oxide, only in the presence of P. alliacea alliinase and its natural substrate, S-benzyl-l-cysteine sulfoxide (petiveriin). Depending on its concentration relative to that of P. alliacea alliinase, the LFS sequesters, to varying degrees, the sulfenic acid intermediate formed by alliinase-mediated breakdown of petiveriin. At LFS:alliinase of 5:1, LFS sequesters all of the sulfenic acid formed by alliinase action on petiveriin, and converts it entirely to (Z)-phenylmethanethial S-oxide. However, starting at LFS:alliinase of 5:2, the LFS is unable to sequester all of the sulfenic acid produced by the alliinase, with the result that sulfenic acid that escapes the action of the LFS condenses with loss of water to form S-benzyl phenylmethanethiosulfinate (petivericin). The results show that the LFS and alliinase function in tandem, with the alliinase furnishing the sulfenic acid substrate on which the LFS acts. The results also show that the LFS modulates the formation of biologically active thiosulfinates that are downstream of the alliinase in a manner dependent upon the relative concentrations of the LFS and the alliinase. These observations suggest that manipulation of LFS-to-alliinase ratios in plants displaying this system may provide a means by which to rationally modify organosulfur small molecule profiles to obtain desired flavor and/or odor signatures, or increase the presence of desirable biologically active small molecules.Lachrymatory factor synthase (LFS) is the term coined to refer to the recently discovered enzyme shown to catalyze the formation of the sulfine responsible for the lachrymatory effect of onion (Allium cepa), (Z)-propanethial S-oxide (PTSO; Imai et al., 2002). Until the discovery of the onion LFS, the formation of the onion lachrymatory factor (LF) was thought to be mediated by only a single enzyme, onion alliinase. Alliinases, which are pyridoxal 5′-P (PLP)-dependent Cys sulfoxide lyases most often found in members of the Allium genus, catalyze the breakdown of Cys sulfoxide derivatives to yield fleeting sulfenic acid intermediates and α-aminoacrylic acid (Scheme 1; Block, 1992; Shimon et al., 2007). Once formed, the sulfenic acids are most often observed to spontaneously condense with loss of water to form thiosulfinates, whereas the α-aminoacrylic acid is further hydrolyzed with loss of ammonia to form pyruvate. The S-substituted Cys sulfoxides that are acted upon by alliinases differ from one another by the identity of the sulfur-bound R group. In Allium plants, the R groups are alk(en)yl, with R = methyl and 2-propenyl appearing in large quantities in garlic (Allium sativum) and R = methyl and (E)-1-propenyl preponderating in onion (Scheme 1). The Cys sulfoxide that serves as the precursor of the onion lachrymator is (E)-S-(1-propenyl)-l-Cys sulfoxide (isoalliin). It is structurally distinct from other naturally occurring S-substituted Cys sulfoxides so far reported in that it is α,β-unsaturated. This structural feature affords its corresponding 1-propenylsulfenic acid (PSA) the possibility of undergoing a [1,4]-sigmatropic rearrangement that, in principle, would furnish the onion lachrymator, PTSO. Indeed, the formation of the onion lachrymator was proposed to occur by such a mechanism (Scheme 2; Block, 1992). Thus, it was surmised that were the α,β-unsaturation to be absent in the precursor S-substituted Cys sulfoxide, the [1,4]-sigmatropic rearrangement that would lead to sulfine formation could not occur. Consequently, it was not surprising that other S-substituted Cys sulfoxides constitutively present in garlic, onion, and other alliinase-containing plants, but devoid of this α,β-unsaturation in the sulfur-bound R group, did not themselves yield lachrymators on plant tissue wounding. It has since been discovered, however, that formation of the onion lachrymator is not catalyzed by onion alliinase, but instead by a novel class of enzyme—LFS. Imai et al. (2002) observed that although a crude preparation of onion alliinase yielded both the LF and the corresponding thiosulfinate, the protein fraction with lachrymator-forming ability could be completely separated from that with alliinase activity by passing the crude onion protein preparation through a hydroxyapatite column. The LFS was subsequently purified and shown to be highly substrate specific, producing the LF from only (E)-S-(1-propenyl)-l-Cys sulfoxide (isoalliin), which occurs constitutively in onion. Interestingly, the LF was detected only when three components, namely, the purified onion alliinase, isoalliin, and the onion LFS, were present in the reaction mixture simultaneously (Imai et al., 2002). Omission of the LFS from the reaction mixture resulted in an increased yield of thiosulfinates, but no LF. Although the complete cDNA sequence of the onion LFS has been determined (Imai et al., 2002), to our knowledge, full biochemical characterization of the enzyme has yet to be reported.Open in a separate windowScheme 1.Alliinase-mediated formation of thiosulfinates from Cys sulfoxide precursors (Block, 1992; Shimon et al., 2007). Alliin is S-allyl-l-Cys sulfoxide, isoalliin is (E)-S-(1-propenyl)-l-Cys sulfoxide, methiin is S-methyl-l-Cys sulfoxide, and propiin is S-propyl-l-Cys sulfoxide.Open in a separate windowScheme 2.Mechanism advanced by Block (1992) to account for formation of the onion lachrymator, PTSO. Alliinase-bound PLP forms a Schiff base with bound isoalliin. General base catalysis at the active site yields an α,β-unsaturated sulfenic acid that can undergo a [1,4]-sigmatropic rearrangement to furnish the sulfine.In the course of our studies on the organosulfur chemistry of non-Allium plants, we isolated and characterized the S-benzyl-l-Cys sulfoxides (petiveriins) and S-(2-hydroxyethyl)-l-Cys sulfoxides (2-hydroxyethiins) from the Amazonian medicinal plant Petiveria alliacea (Fig. 1; Kubec and Musah, 2001; Kubec et al., 2002). These compounds are S-substituted Cys sulfoxide derivatives with R = benzyl and 2-hydroxyethyl, respectively, that, to our knowledge, had never before been isolated from plants. We showed that, as has been observed in garlic and onion, symmetrical and mixed thiosulfinate derivatives of the corresponding petiveriin and 2-hydroxyethiin precursors could be extracted with ether solvent (Fig. 1; Kubec et al., 2002) upon root tissue disruption. We have also shown that an alliinase that mediates the transformation of the petiveriins and 2-hydroxyethiins to their corresponding thiosulfinates is present in P. alliacea (Musah et al., 2009). Interestingly, while working with P. alliacea root extracts, we noted the presence of a potent lachrymator that we subsequently determined to be a sulfine—(Z)-phenylmethanethial S-oxide (PMTSO; Fig. 2; Kubec et al., 2003). However, the biochemical precursor of PMTSO and the pathway(s) leading to its formation upon disruption of P. alliacea tissue remain to be determined. Given that the onion LF (PTSO), whose formation is mediated by an LFS, is also a sulfine, we were prompted to investigate the possibility of the presence of a LFS in P. alliacea. In this report, we describe our confirmation of the existence of a LFS in P. alliacea, and detail biochemical characterization of this novel class of enzymes.Open in a separate windowFigure 1.Cys sulfoxides and their corresponding thiosulfinate derivatives isolated from the Amazonian medicinal plant P. alliacea. The breakdown of the Cys sulfoxides is mediated by P. alliacea alliinase.Open in a separate windowFigure 2.Lachrymatory sulfine isolated from P. alliacea.  相似文献   

18.
19.
The oxidative pentose phosphate pathway is required for function of the alternative pyrimidine biosynthetic pathway, a pathway that allows thiamine synthesis in the absence of the PurF enzyme in Salmonella typhimurium. Mutants that no longer required function of the oxidative pentose phosphate pathway for thiamine synthesis were isolated. Further phenotypic analyses of these mutants demonstrated that they were also sensitive to the presence of serine in the medium, suggesting a partial defect in isoleucine biosynthesis. Genetic characterization showed that these pleiotropic phenotypes were caused by null mutations in yjgF, a previously uncharacterized open reading frame encoding a hypothetical 13.5-kDa protein. The YjgF protein belongs to a class of proteins of unknown function that exhibit striking conservation across a wide range of organisms, from bacteria to humans. This work represents the first detailed phenotypic characterization of yjgF mutants in any organism and provides important clues as to the function of this highly conserved class of proteins. Results also suggest a connection between function of the isoleucine biosynthetic pathway and the requirement for the pentose phosphate pathway in thiamine synthesis.The increasing number of completed genome sequences has resulted in the identification of new families of hypothetical proteins whose function has yet to be established. The lack of existing mutants defective in these conserved proteins suggests novel, complex, or subtle phenotypes. Through our work on thiamine synthesis in Salmonella typhimurium, we have isolated mutants defective in the recently identified YER057c/YjgF protein family. Our data suggest that defects in this protein result in complex phenotypes involving thiamine and isoleucine biosynthesis.Thiamine pyrophosphate (TPP) serves as an essential cofactor for a number of metabolic reactions involving the removal or transfer of C2 units. Despite the important role of TPP in cellular metabolism, its synthesis and regulation are not well understood in any organism. TPP is formed from two precursors, 4-methyl-5-(β-hydroxyethyl)thiazole phosphate (THZ-P) and 4-amino-5-hydroxymethyl-2-methylpyrimidine pyrophosphate (HMP-PP). These compounds are joined and subsequently phosphorylated as shown in Fig. Fig.1A.1A. Although many of the enzymatic steps in both the THZ-P and HMP-PP pathways have not been clearly defined, the major precursor molecules for both of these compounds have been determined by labeling studies (17, 20, 28, 29). In particular, the purine pathway intermediate, aminoimidazole ribotide (AIR), has been shown to provide all of the atoms in HMP (28, 50, 51).Open in a separate windowFIG. 1Pathway schematics. (A) Biosynthetic pathway for TPP. The involvement of the purine pathway in HMP-PP synthesis is shown with structural intermediates prior to the AIR branch point. Arrows denoted with dotted lines represent proposed steps. Reactions involved in the conversion of AIR to HMP-PP and in the synthesis of THZ-P have not been clearly defined. Genes whose products are required for selected reactions are indicated next to the relevant arrows. Abbreviations: R-P, ribose-5-phosphate, PRPP, phosphoribosylpyrophosphate. (B) Biosynthetic pathways for the branched-chain amino acids isoleucine and valine. Enzymes that catalyze specific steps are as follows: 1, aspartate transaminase; 2, 3, and 4, aspartate kinases I, II, and III, respectively; 5, aspartate semialdehyde dehydrogenase; 6 and 7, homoserine dehydrogenases I and II, respectively; 8, homoserine kinase; 9, threonine synthase; 10, threonine deaminase; 11 and 12, acetohydroxy acid synthases I and II, respectively; 13, acetohydroxy acid isomeroreductase; 14, dihydroxy acid dehydratase; 15, transaminase B; 16, transaminase C. OAA, oxaloacetic acid.Although the involvement of the purine pathway in the synthesis of HMP is clear, there is substantial genetic and biochemical evidence indicating that the first enzyme of the purine pathway, phosphoribosylpyrophosphate amidotransferase (PurF) (EC 2.4.2.14), is not required for HMP synthesis in S. typhimurium under all conditions. Mutants defective in purF are able to grow in the absence of thiamine when glucose is used as a carbon source if pantothenate is also supplied in the medium (23). Similarly, purF mutants do not require thiamine when grown on a number of nonglucose carbon sources, such as gluconate or ribose (54). The pathway responsible for synthesis of HMP independent of the PurF enzyme has been defined as the alternative pyrimidine biosynthetic (APB) pathway (21, 54); recent biochemical data suggest that phosphoribosylamine (PRA), or a derivative, is an intermediate in this pathway (24).Significant progress in our understanding of the APB pathway has been made by the isolation and characterization of mutants unable to synthesize thiamine in a purF background. One class of mutants, designated apbA, was defective in a pantothenate biosynthetic enzyme (ketopantoate reductase [PanE]) (32, 33), consistent with previous results implicating a role for pantothenate in PurF-independent thiamine synthesis (23). A second class of these mutants was defective in the oxidative pentose phosphate pathway, affecting either glucose-6-phosphate dehydrogenase (Zwf) or gluconate-6-phosphate dehydrogenase (Gnd) (25, 54). Addition of ribose-5-phosphate (ribose-5-P) restored function of the APB pathway in these mutants, suggesting that the role of these enzymes in HMP synthesis was to supply ribose-5-P. These results led to the model shown in Fig. Fig.1A1A which implicates ribose-5-P and an amine donor as precursors to PRA. Repeated attempts have failed to identify either the predicted PRA-forming activity or mutants defective in this step (27). There are several possible explanations for this. It is possible that the correct substrates have not been identified and/or that the PRA-forming activity is required for another cellular function.In this report, we describe the isolation and characterization of mutations that allow function of the APB pathway in the absence of the pentose phosphate pathway. These mutations were found to disrupt a previously uncharacterized open reading frame (ORF) encoding a hypothetical 13.5-kDa protein. We have designated this gene yjgF based on homology to the respective ORF in Escherichia coli. The YjgF protein belongs to the YER057c/YjgF protein family, a class of proteins of unknown function that exhibit striking conservation across a wide range of organisms. Characterization of these mutants revealed that they also were sensitive to the presence of serine in the medium, exhibiting a requirement for isoleucine under this condition. The phenotypes caused by yjgF mutations suggest that the YjgF protein may be involved in regulation or function of the isoleucine biosynthetic pathway. Further, results suggest a connection between isoleucine biosynthesis and function of the APB pathway in thiamine synthesis.  相似文献   

20.
The lantibiotic nisin is a potent antimicrobial substance, which contains unusual lanthionine rings and dehydrated amino acid residues and is produced by Lactococcus lactis. Recently, the nisin biosynthetic machinery has been applied to introduce lanthionine rings in peptides other than nisin with potential therapeutic use. Due to difficulties in the isolation of the proposed synthetase complex (NisBTC), mechanistic information concerning the enzymatic biosynthesis of nisin is scarce. Here, we present the molecular characterization of a number of nisin mutants that affect ring formation. We have investigated in a systematic manner how these mutations influence dehydration events, which are performed enzymatically by the dehydratase NisB. Specific mutations that hampered ring formation allowed for the dehydration of serine residues that directly follow the rings and are normally unmodified. The combined information leads to the conclusion that 1) nisin biosynthesis is an organized directional process that starts at the N terminus of the molecule and continues toward the C terminus, and 2) NisB and NisC are alternating enzymes, whose activities follow one after another in a repetitive way. Thus, the dehydration and cyclization processes are not separated in time and space. On the basis of these results and previous knowledge, a working model for the sequence of events in the maturation of nisin is proposed.Nisin is a lantibiotic produced by Lactococcus lactis, which has been known since 1928 (1, 2). This antimicrobial peptide is active against various Gram-positive bacteria and has attained commercial success as a food preservative (3). In addition to the wide industrial applications of nisin, it became also a model system to study various aspects of lantibiotic biosynthesis, regulation, and mode of action (2). Furthermore, recently, other applications of nisin have emerged. Its biosynthetic machinery can be successfully used to install dehydrated amino acids and lanthionine rings in peptides, which are either related or totally unrelated to nisin (411). This offers great opportunities to modulate the stability and activity of peptides that are used as therapeutics (8).The post-translational modified nisin molecule is classified as a member of the Group A lantibiotics (12). Mature nisin contains 34 amino acids, three of which are posttranslationally modified, and five thioether rings that are enzymatically formed upon cyclization of five free cysteines and five dehydroamino acid residues (Fig. 1). These peculiar modifications, which are very rare in nature, give nisin its exceptional stability against proteolysis and contribute greatly to its antimicrobial activity.Open in a separate windowFIGURE 1.Primary structure of prenisin and generated mutants. Dehydrated residues are shaded gray; serine 33 sometimes escapes dehydration and is shaded light gray. Serine at position 29 is never dehydrated in wild type prenisin. The impact of mutations on the dehydration pattern of new prenisin species is schematically depicted. Mutated residues are indicated by filled red circles. Newly formed dehydrated residues are pointed to by a black arrow. Letters A–E correspond to the five consecutive lanthionine rings in nisin.Nisin is synthesized ribosomally as a 57-amino acid residue-long polypeptide. Subsequently, it is directed to a putative synthetase complex that probably consists of three different proteins that include the dehydratase NisB, responsible for dehydration of serines and threonines to dehydroalanines and dehydrobutyrines, respectively; the cyclase NisC, which forms (methyl) lanthionine bridges between cysteines and dehydroamino acids; and the ABC transporter NisT, which performs transport across the lipid bilayer by consuming ATP. Newly synthesized and modified prenisin is still antimicrobially inactive. Only upon cleavage of the leader sequence that encompasses the first 23 amino acids by the dedicated protease NisP, an active molecule is liberated.Although there are data pointing to the existence of a synthetase complex that modifies nisin, such a complex has not been isolated so far. However, both NisB (13, 14) and NisC (13) were shown by specific antibody detection to localize at the cytoplasmic membrane, although some soluble signal was also detected. This localization gives NisBC the opportunity to interact with the transporter NisT, which is an integral membrane protein. Furthermore, co-immunoprecipitation and yeast two-hybrid studies suggested an interaction between members of the nisin modification machinery and nisin itself (13). The function of each member of the putative multimeric synthetase has been investigated in vivo by knock-out studies. It also has been demonstrated that subsequent steps in nisin biosynthesis can be performed separately. Dehydration, cyclization, and transport of the modified product were dissected in vivo, and also the dehydratase has been shown to perform enzymatic reactions without the presence of other members of the complex in vivo (7) although with very low efficiency. The cyclization activity of NisC was demonstrated in vitro (15), and the ABC transporter NisT was shown to be capable of transport of unmodified prenisin in vivo (10). Based on the available data, it is difficult to assess whether multimeric lanthionine complexes are indispensable for efficient nisin production and modification. However, in vivo localization studies and interaction experiments suggest that these proteins work in a concerted manner.Here, we present data that indicates a strong coordination between members of the nisin modification machinery. The analysis of sets of nisin mutants, where key residues that take part in ring formation as well as substitutions of residues that directly follow lanthionine structures, suggests a strong interdependency of dehydratase and cyclase activity. Moreover, the data indicate that these enzymes alternate during catalysis and that they are intertwined in time and space. Our data also suggest that nisin modification is an ordered process that proceeds consecutively from the N terminus of prenisin toward its C terminus. Based on the available literature data and the data presented here, we propose a model wherein nisin is being posttranslationally modified in consecutive steps from its N terminus toward its C terminus.  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