首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.

Introduction

Molecular factors are differentially observed in various bent sectors of poplar (Populus nigra) woody taproots. Responses to stress are modulated by a complex interplay among different hormones and signal transduction pathways. In recent years, metabolomics has been recognized as a powerful tool to characterize metabolic network regulation, and it has been widely applied to investigate plant responses to biotic and abiotic stresses.

Objectives

In this paper we used metabolomics to understand if long term-bending stress induces a “spatial” and a “temporal” metabolic reprogramming in woody poplar roots.

Methods

By NMR spectroscopy and statistical analysis we investigated the unstressed and three portions of stressed root (above-bent, bent, and below-bent) sectors collected at 12 (T0), 13 (T1) and 14 (T2) months after stress induction.

Results

The data indicate a clear between-class separation of control and stressed regions, based on the metabolites regulation, during both spatial and temporal changes. We found that taproots, as a consequence of the stress, try to restore homeostasis and normal metabolic fluxes thorough the synthesis and/or accumulation of specific compounds related to mechanical forces distribution along the bent taproot.

Conclusion

The data demonstrate that the impact of mechanical stress on plant biology can efficiently be studied by NMR-based metabolomics.
  相似文献   

2.
In the seeds of Hevea brasiliensis, the cyanogenic monoglucoside linamarin (2-β-d-glucopyranosyloxy-2-methylpropionitrile) is accumulated in the endosperm. After onset of germination, the cyanogenic diglucoside linustatin (2-[6-β-d-glucosyl-β-d-glucopyranosyloxy]-2- methylpropionitrile) is formed and exuded from the endosperm of Hevea seedlings. At the same time the content of cyanogenic monoglucosides decreases. The linustatin-splitting diglucosidase and the β-cyanoalanine synthase that assimilates HCN, exhibit their highest activities in the young seedling at this time. Based on these observations the following pathway for the in vivo mobilization and metabolism of cyanogenic glucosides is proposed: storage of monoglucosides (in the endosperm)—glucosylation—transport of the diglucoside (out of the endosperm into the seedling)—cleavage by diglucosidase—reassimilation of HCN to noncyanogenic compounds. The presence of this pathway demonstrates that cyanogenic glucosides, typical secondary plant products serve in the metabolism of developing plants as N-storage compounds and do not exclusively exhibit protective functions due to their repellent effect.  相似文献   

3.

Background

Phytophthora infestans is a plant pathogen that causes an important plant disease known as late blight in potato plants (Solanum tuberosum) and several other solanaceous hosts. This disease is the main factor affecting potato crop production worldwide. In spite of the importance of the disease, the molecular mechanisms underlying the compatibility between the pathogen and its hosts are still unknown.

Results

To explain the metabolic response of late blight, specifically photosynthesis inhibition in infected plants, we reconstructed a genome-scale metabolic network of the S. tuberosum leaf, PstM1. This metabolic network simulates the effect of this disease in the leaf metabolism. PstM1 accounts for 2751 genes, 1113 metabolic functions, 1773 gene-protein-reaction associations and 1938 metabolites involved in 2072 reactions. The optimization of the model for biomass synthesis maximization in three infection time points suggested a suppression of the photosynthetic capacity related to the decrease of metabolic flux in light reactions and carbon fixation reactions. In addition, a variation pattern in the flux of carboxylation to oxygenation reactions catalyzed by RuBisCO was also identified, likely to be associated to a defense response in the compatible interaction between P. infestans and S. tuberosum.

Conclusions

In this work, we introduced simultaneously the first metabolic network of S. tuberosum and the first genome-scale metabolic model of the compatible interaction of a plant with P. infestans.
  相似文献   

4.
In the present investigation, metabolites of Streptomyces sp. MTN14 and Trichoderma harzianum ThU significantly enhanced biomass yield (3.58 and 3.48 fold respectively) in comparison to the control plants. The secondary metabolites treatments also showed significant augmentation (0.75–2.25 fold) in withanolide A, a plant secondary metabolite. Lignin deposition, total phenolic and flavonoid content in W. somnifera were maximally induced in treatment having T. harzianum metabolites. Also, Trichoderma and Streptomyces metabolites were found much better in invoking in planta contents and antioxidants compared with their live culture treatments. Therefore, identification of new molecular effectors from metabolites of efficient microbes may be used as biopesticide and biofertilizer for commercial production of W. somnifera globally.  相似文献   

5.
The endosperm of castor bean seeds (Ricinus communis L.) contains two —SH-dependent aminopeptidases, one hydrolyzing l-leucine-β-naphthylamide optimally at pH 7.0, and the other hydrolyzing l-proline-β-naphthylamide optimally at pH 7.5. After germination the endosperm contains in addition an —SH-dependent hemoglobin protease, a serine-dependent carboxypeptidase, and at least two —SH-dependent enzymes hydrolyzing the model substrate α-N-benzoyl-dl-arginine-β-naphthylamide (BANA). The carboxypeptidase is active on a variety of N-carbobenzoxy dipeptides, especially N-carbobenzoxy-L-phenylalanine-l-alanine and N-carbobenzoxy-l-tyrosine-l-leucine. The pH optima for the protease, carboxypeptidase, and BANAase acivities are 3.5 to 4.0, 5.0 to 5.5, and 6 to 8, respectively.  相似文献   

6.

Main conclusion

Medicinal and aromatic plants are known to produce secondary metabolites that find uses as flavoring agents, fragrances, insecticides, dyes and drugs. Biotechnology offers several choices through which secondary metabolism in medicinal plants can be altered in innovative ways, to overproduce phytochemicals of interest, to reduce the content of toxic compounds or even to produce novel chemicals. Detailed investigation of chromatin organization and microRNAs affecting biosynthesis of secondary metabolites as well as exploring cryptic biosynthetic clusters and synthetic biology options, may provide additional ways to harness this resource. Plant secondary metabolites are a fascinating class of phytochemicals exhibiting immense chemical diversity. Considerable enigma regarding their natural biological functions and the vast array of pharmacological activities, amongst other uses, make secondary metabolites interesting and important candidates for research. Here, we present an update on changing trends in the biotechnological approaches that are used to understand and exploit the secondary metabolism in medicinal and aromatic plants. Bioprocessing in the form of suspension culture, organ culture or transformed hairy roots has been successful in scaling up secondary metabolite production in many cases. Pathway elucidation and metabolic engineering have been useful to get enhanced yield of the metabolite of interest; or, for producing novel metabolites. Heterologous expression of putative plant secondary metabolite biosynthesis genes in a microbe is useful to validate their functions, and in some cases, also, to produce plant metabolites in microbes. Endophytes, the microbes that normally colonize plant tissues, may also produce the phytochemicals produced by the host plant. The review also provides perspectives on future research in the field.
  相似文献   

7.
Laser-ablation electrospray ionization (LAESI)-mass spectrometry imaging has been applied to contrasting plant organs to assess its potential as a procedure for performing in vivo metabolomics in plants. In a proof-of-concept experiment, purple/white segmented Phalaenopsis spp. petals were first analyzed using standard liquid chromatography-mass spectrometry analyses of separate extracts made specifically from the purple and white regions. Discriminatory compounds were defined and putatively annotated. LAESI analyses were then performed on living tissues, and these metabolites were then relocalized within the LAESI-generated data sets of similar tissues. Maps were made to illustrate their locations across the petals. Results revealed that, as expected, anthocyanins always mapped to the purple regions. Certain other (nonvisible) polyphenols were observed to colocalize with the anthocyanins, whereas others were found specifically within the white tissues. In a contrasting example, control and Cladosporium fulvum-infected tomato (Solanum lycopersicum) leaves were subjected to the same procedures, and it could be observed that the alkaloid tomatine has clear heterogeneous distribution across the tomato leaf lamina. Furthermore, LAESI analyses revealed perturbations in alkaloid content following pathogen infection. These results show the clear potential of LAESI-based imaging approaches as a convenient and rapid way to perform metabolomics analyses on living tissues. However, a range of limitations and factors have also been identified that must be taken into consideration when interpreting LAESI-derived data. Such aspects deserve further evaluation before this approach can be applied in a routine manner.Plants are a tremendously rich source of a myriad of structurally and chemically diverse metabolites (Rao and Ravishankar, 2002; D’Auria and Gershenzon, 2005). Many of these metabolites have a (partly) known function in the plant, although our knowledge of the vast majority of plant secondary metabolites is still sparse, or even nonexistent (Rao and Ravishankar, 2002; D’Auria and Gershenzon, 2005; Fernie, 2007). Plant metabolites are also of considerable importance in a crop context. Indeed, most plant species that have undergone domestication have become crops specifically because they provide us with a source of chemicals. This is not only true for all of our food crops, but also for many other species of genera such as Pyrethrum (insecticides), Jasminium and Santalum (perfumes), Hevea (rubber), Nicotiana and Cannabis (drugs), Linum (oils), Artemisia and Taxus (pharmaceuticals), Cinnamomum (flavors), etc. However, despite the importance of plants as a source of exploitable and essential biochemicals, we often still have remarkably limited knowledge of the relevant biosynthetic pathways, the genetics behind the key enzymes, and indeed when, why, and where these metabolites are produced and stored within the plant in question (Fernie, 2007; Sumner et al., 2011; Kueger et al., 2012).The field of plant metabolomics has grown tremendously since its recent inception earlier this century (Fiehn et al., 2000; Fiehn, 2002). As an untargeted approach to gain a broad overview of the complexity of plant metabolic composition, the technology has, in a short time, made significant inroads into helping expand our knowledge of plant biochemistry (Kueger et al., 2012; Etalo et al., 2013; Hunerdosse and Nomura, 2014; Meret et al., 2014). Typically, rich metabolomics data sets already provide us with a valuable means to generate hypotheses relating to plant metabolism, which then become the focus of further, more direct investigation (Quanbeck et al., 2012). New technologies are being developed, and especially, new data-mining strategies are being designed to allow us to look deep into plant metabolism without having first to rely on preconceptions. However, there are significant limitations to the application of the technology, which still remain the topic of much research effort.Robust sampling approaches for plant biochemical analysis generally entail taking reliably measurable amounts of plant material that will yield detectable levels of the chemical components. Although for metabolomics analyses, samples of just 50 mg can often suffice, obtaining a reliable sample with minimum biological variation generally requires an initial pooling of materials from which a representative sample is then taken. We therefore treat plant tissue as being homogeneous, but this is clearly a gross oversimplification (Fernie, 2007). Plants have been considered to be composed of roughly 40 different cell types, and a plant organ such as a leaf will generally contain up to 15 different cell types (Martin et al., 2001). Different morphologies also parallel different biochemical composition. Even directly neighboring cells within an organ, for example, a leaf epidermis that often comprises pavement, guard, trichome, and glandular hair cells, are formed from cells already known to have distinctly different biochemistries. Making an extract, for any kind of metabolomics or standard biochemical analysis, therefore entails that we immediately lose most intercellular and intertissue resolution. However, our knowledge is growing in that, in addition to known or expected biochemical differences between cell types, metabolite accumulation across organs can be far from uniform; indeed, islands of higher and lower concentrations of particular metabolites have been observed. This is of course immediately visible when the metabolites concerned can be seen by the naked eye; anthocyanins, for example, are often found to be heterogeneously distributed across leaves, fruits, and flower petals, creating clear phenotypic patterns. The same may also be true of other compounds that are invisible to the human eye but that, in contrast, may still be detectable by insects (e.g. through their fluorescence capacity; see http://www.naturfotograf.com/UV_flowers_list.html; Gronquist et al., 2001).In an ideal situation, we would like to be able to look directly into a plant tissue and be able to analyze the biochemical composition at the single cell level. Some so-called metabolite imaging technologies, usually based on mass spectrometric detection (mass spectrometry imaging [MSI]), have recently been introduced as a step toward this optimistic goal. Included here are matrix-assisted laser desorption/ionization (MALDI)-MSI, direct analysis in real time, and desorption electrospray ionization approaches (Cody et al., 2005; Cornett et al., 2007; Ifa et al., 2010). Early examples of MALDI-MSI have shown not only how primary metabolites such as sugars can be strongly localized within plant organs (Rolletschek et al., 2011), but also how the heterogeneous distribution of glucosinolates in Arabidopsis (Arabidopsis thaliana) can potentially determine grazing behavior of caterpillars (Shroff et al., 2008). This technology continues to improve, and recent exciting developments have resulted in cellular and subcellular imaging of metabolites at a resolution of 5 to 9 µm using MALDI (Korte et al., 2015). However, some key practical limitations of MALDI-based approaches are centered around the need to initially have to pretreat/dehydrate the tissue prior to applying the required matrix solution and the requirement of applying a vacuum during the biochemical analysis. Recently, a new technology has been introduced, laser ablation electrospray ionization (LAESI), which can potentially overcome some of these limitations, given that measurements can be made on fresh, living tissue without the need for a vacuum, thus creating the potential for high-resolution in vivo metabolomics.Here, we report on a set of experiments performed to assess both the potential and limitations of using LAESI-based MSI approaches to perform metabolic mapping on living plant tissues. While identifying a number of technological challenges that still need to be tackled, we were able to show that it is possible to use LAESI-MSI to map metabolites directly onto their known location (in this case, by exploiting the visibility of anthocyanins) as well as localize invisible metabolites in the same tissue. Results have revealed that in plants, for both petal and leaf tissue, the distribution of metabolites can be highly heterogeneous, and that this heterogeneity is of potential relevance to our gaining a broader, more detailed understanding of the overall molecular organization and phenotypic features of plant tissues. Furthermore, knowledge of the nature and extent of this heterogeneity has particular relevance and importance when trying to understand how a plant functions as a system, interacting with its environment. We predict that a higher resolution understanding of plant biochemistry will lead to an increasingly discriminatory capacity in our ability to define more accurately the spatial complexity of plant molecular organization.  相似文献   

8.
Deciphering the influence of genetics on primary metabolism in plants will provide insights useful for genetic improvement and enhance our fundamental understanding of plant growth and development. Although maize (Zea mays) is a major crop for food and feed worldwide, the genetic architecture of its primary metabolism is largely unknown. Here, we use high-density linkage mapping to dissect large-scale metabolic traits measured in three different tissues (leaf at seedling stage, leaf at reproductive stage, and kernel at 15 d after pollination [DAP]) of a maize recombinant inbred line population. We identify 297 quantitative trait loci (QTLs) with moderate (86.2% of the mapped QTL, R2 = 2.4 to 15%) to major effects (13.8% of the mapped QTL, R2 >15%) for 79 primary metabolites across three tissues. Pairwise epistatic interactions between these identified loci are detected for more than 25.9% metabolites explaining 6.6% of the phenotypic variance on average (ranging between 1.7 and 16.6%), which implies that epistasis may play an important role for some metabolites. Key candidate genes are highlighted and mapped to carbohydrate metabolism, the tricarboxylic acid cycle, and several important amino acid biosynthetic and catabolic pathways, with two of them being further validated using candidate gene association and expression profiling analysis. Our results reveal a metabolite-metabolite-agronomic trait network that, together with the genetic determinants of maize primary metabolism identified herein, promotes efficient utilization of metabolites in maize improvement.  相似文献   

9.

Introduction

Usually whole plant or whole leaf extracts are analyzed to study the chemical ecology of insect-plant interactions. For herbivore species the contact with the leaf surface enables them to estimate the quality of the plant. The relationship between the leaf-surface and leaf-tissue secondary metabolites (SMs) could offer important new insights in insect-plant interactions mediated by SMs. Pyrrolizidine alkaloids (PAs), typical defense chemicals in Jacobaea species, are repellent for generalist herbivores but are attractive to specialists.

Objectives

Explore whether the PAs on the leaf surface are a reliable representation of the PAs in the leaf tissue in PA-containing plants.

Method

The concentration of individual PAs present on the leaf surface and in the corresponding leaf tissue from 37 genotypes (one plant from each genotype) of an F2 generation of a cross between Jacobaea vulgaris and Jacobaea aquatica was measured by high performance liquid chromatography coupled to tandem mass spectrometry (LC-MS/MS). PAs were removed from the leaf surface by extraction with a slightly acidic aqueous solution.

Results

The total amount of PAs present on the surface of the leaves was only 0.015% (range 0.001–0.163%) of the total amount present in the leaf tissue. Most PAs present in the leaf tissue were also found on the surface, except for jaconine, dehydrojaconine, dehydrojacoline and usaramine N-oxide. Positive correlations between leaf-surface and leaf-tissue concentrations were found for most of the jacobine-like and otosenine-like PAs, but correlations for total PA, senecionine- and erucifoline-like PAs were not significant.

Conclusion

These results indicate that PA variation on the leaf surface only partially reflects the PA variation in the leaf tissue. Because most herbivores are affected in a different manner by individual PAs, this result means that the leaf surface does not give a reliable estimate of plant quality to herbivores.
  相似文献   

10.

Background

Plant secondary metabolites act as defence molecules to protect plants from biotic and abiotic stresses. In particular, C-glycosylated flavonoids are more stable and reactive than their O-glycosylated counterparts. Therefore, vitexin (apigenin 8-C glucoside) present in Ficus deltoidea is well-known for its antioxidant, anti-inflammatory, and antidiabetic properties.

Methods

Phenol based extraction was used to extract proteins (0.05% yield) with less plant pigments. This can be seen from clear protein bands in gel electrophoresis. In-gel trypsin digestion was subsequently carried out and analysed for the presence of peptides by LC-MS/MS.

Results

Thirteen intact proteins are identified on a 12% polyacrylamide gel. The mass spectra matching was found to have 229 proteins, and 11.4% of these were involved in secondary metabolism. Proteins closely related to vitexin biosynthesis are listed and their functions are explained mechanistically. Vitexin synthesis is predicted to involve plant polyketide chalcone synthase, isomerization by chalcone isomerase, oxidation by cytochrome P450 to convert flavanone to flavone, and transfer of sugar moiety by C-glycosyltransferase, followed by dehydration to produce flavone-8-C-glucosides.

Conclusions

Phenol based extraction, followed by gel electrophoresis and LC-MS/MS could identify proteome explaining vitexin biosynthesis in F. deltoidea. Many transferases including β-1,3-galactosyltransferase 2 and glycosyl hydrolase family 10 protein were detected in this study. This explains the importance of transferase family proteins in Cglycosylated apigenin biosynthesis in medicinal plant.
  相似文献   

11.

Introduction

Onion (Allium cepa) represents one of the most important horticultural crops and is used as food, spice and medicinal plant almost worldwide. Onion bulbs accumulate a broad range of primary and secondary metabolites which impact nutritional, sensory and technological properties.

Objectives

To complement existing analytical methods targeting individual compound classes this work aimed at the development and validation of an analytical workflow for comprehensive metabolite profiling of onion bulbs.

Method

Metabolite profiling was performed by liquid chromatography coupled with electrospray ionization quadrupole time-of-flight mass spectrometry (LC/ESI-QTOFMS). For annotation of metabolites accurate mass tandem mass spectrometry experiments were carried out.

Results

On the basis of LC/ESI-QTOFMS and two chromatographic methods an analytical workflow was developed which facilitates profiling of polar and semi-polar onion metabolites including fructooligosaccharides, proteinogenic amino acids, peptides, S-substituted cysteine conjugates, flavonoids and saponins. To minimize enzymatic conversion of S-alk(en)ylcysteine sulfoxides, a sample preparation and extraction protocol for fresh onions was developed comprising cryohomogenization and a low-temperature quenching step. A total of 123 metabolites were annotated and characterized by chromatographic and tandem mass spectral data. For validation, recovery rates and matrix effects were determined for 15 model compounds. Repeatability and linearity were assessed for more than 80 endogenous metabolites.

Conclusion

As exemplarily demonstrated by comparative metabolic analysis of six onion cultivars the established analytical workflow in combination with targeted and non-targeted data analysis strategies can be successfully applied for comprehensive metabolite profiling of onion bulbs.
  相似文献   

12.
White rot fungi can oxidize high-molecular-weight polycyclic aromatic hydrocarbons (PAH) rapidly to polar metabolites, but only limited mineralization takes place. The objectives of this study were to determine if the polar metabolites can be readily mineralized by indigenous microflora from several inoculum sources, such as activated sludge, forest soils, and PAH-adapted sediment sludge, and to determine if such metabolites have decreased mutagenicity compared to the mutagenicity of the parent PAH. 14C-radiolabeled benzo[a]pyrene was subjected to oxidation by the white rot fungus Bjerkandera sp. strain BOS55. After 15 days, up to 8.5% of the [14C]benzo[a]pyrene was recovered as 14CO2 in fungal cultures, up to 73% was recovered as water-soluble metabolites, and only 4% remained soluble in dibutyl ether. Thin-layer chromatography analysis revealed that many polar fluorescent metabolites accumulated. Addition of indigenous microflora to fungal cultures with oxidized benzo[a]pyrene on day 15 resulted in an initially rapid increase in the level of 14CO2 recovery to a maximal value of 34% by the end of the experiments (>150 days), and the level of water-soluble label decreased to 16% of the initial level. In fungal cultures not inoculated with microflora, the level of 14CO2 recovery increased to 13.5%, while the level of recovery of water-soluble metabolites remained as high as 61%. No large differences in 14CO2 production were observed with several inocula, showing that some polar metabolites of fungal benzo[a]pyrene oxidation were readily degraded by indigenous microorganisms, while other metabolites were not. Of the inocula tested, only PAH-adapted sediment sludge was capable of directly mineralizing intact benzo[a]pyrene, albeit at a lower rate and to a lesser extent than the mineralization observed after combined treatment with white rot fungi and indigenous microflora. Fungal oxidation of benzo[a]pyrene resulted in rapid and almost complete elimination of its high mutagenic potential, as observed in the Salmonella typhimurium revertant test performed with strains TA100 and TA98. Moreover, no direct mutagenic metabolite could be detected during fungal oxidation. The remaining weak mutagenic activity of fungal cultures containing benzo[a]pyrene metabolites towards strain TA98 was further decreased by subsequent incubations with indigenous microflora.Bioremediation of polycyclic aromatic hydrocarbon (PAH)-polluted soil is severely hampered by the low rate of degradation of the higher PAH, particularly the four- and five-ring PAH (6, 32). These higher PAH have very low water solubility and are often tightly bound to soil particles. This results in very low bioavailability for bacterial degradation. The observation that white rot fungi can oxidize PAH rapidly with their extracellular ligninolytic enzyme systems has therefore raised interest in the use of these organisms for bioremediation of PAH-polluted soils (3, 9). Although PAHs are extensively oxidized by white rot fungi, the degree of mineralization to CO2 is always limited. In various studies evaluating the degradation of the potent carcinogen benzo[a]pyrene by several white rot fungal species, from 0.17 to 19% of the radiolabeled PAH was recovered as 14CO2 (4, 5, 26). The major products of the oxidation were both nonpolar and polar metabolites. The accumulation of such metabolites could be a reason for concern, since mammalian and fungal monooxygenases can oxidize benzo[a]pyrene to epoxides and dihydrodiols, which are very potent carcinogens (28, 29). However, peroxidase-mediated extracellular oxidation of benzo[a]pyrene in cultures of white rot fungi results initially in benzo[a]pyrenediones, which show weak mutagenic activity (29). These primary metabolites are rapidly oxidized further to unidentified metabolites by Phanerochaete laevis and Phanerochaete chrysosporium (5, 26). Furthermore, the oxidized benzo[a]pyrene metabolites have a higher aqueous solubility. Since the low bioavailability of PAH is a major rate-limiting factor in the degradation of these compounds by bacteria (27, 31), the increased bioavailability of oxidized PAH metabolites suggests that these compounds can be more easily mineralized by bacteria.The aim of this study was to investigate the degradation and mineralization of the five-ring PAH benzo[a]pyrene by the white rot fungus Bjerkandera sp. strain BOS55 and the subsequent mineralization of the metabolites by natural mixed cultures of microorganisms. During the oxidation and mineralization of benzo[a]pyrene, the decrease in the mutagenicity of the metabolites was monitored. The white rot fungal strain Bjerkandera sp. strain BOS55 was used because of its outstanding ability to rapidly oxidize PAH (8, 19) and because extensive information concerning its physiology is available (7, 18, 20, 22, 23).  相似文献   

13.

Background

Metabolites disrupted by abnormal state of human body are deemed as the effect of diseases. In comparison with the cause of diseases like genes, these markers are easier to be captured for the prevention and diagnosis of metabolic diseases. Currently, a large number of metabolic markers of diseases need to be explored, which drive us to do this work.

Methods

The existing metabolite-disease associations were extracted from Human Metabolome Database (HMDB) using a text mining tool NCBO annotator as priori knowledge. Next we calculated the similarity of a pair-wise metabolites based on the similarity of disease sets of them. Then, all the similarities of metabolite pairs were utilized for constructing a weighted metabolite association network (WMAN). Subsequently, the network was utilized for predicting novel metabolic markers of diseases using random walk.

Results

Totally, 604 metabolites and 228 diseases were extracted from HMDB. From 604 metabolites, 453 metabolites are selected to construct the WMAN, where each metabolite is deemed as a node, and the similarity of two metabolites as the weight of the edge linking them. The performance of the network is validated using the leave one out method. As a result, the high area under the receiver operating characteristic curve (AUC) (0.7048) is achieved. The further case studies for identifying novel metabolites of diabetes mellitus were validated in the recent studies.

Conclusion

In this paper, we presented a novel method for prioritizing metabolite-disease pairs. The superior performance validates its reliability for exploring novel metabolic markers of diseases.
  相似文献   

14.
Holme P 《PloS one》2011,6(2):e16605

Background

Several studies have mentioned network modularity—that a network can easily be decomposed into subgraphs that are densely connected within and weakly connected between each other—as a factor affecting metabolic robustness. In this paper we measure the relation between network modularity and several aspects of robustness directly in a model system of metabolism.

Methodology/Principal Findings

By using a model for generating chemical reaction systems where one can tune the network modularity, we find that robustness increases with modularity for changes in the concentrations of metabolites, whereas it decreases with changes in the expression of enzymes. The same modularity scaling is true for the speed of relaxation after the perturbations.

Conclusions/Significance

Modularity is not a general principle for making metabolism either more or less robust; this question needs to be addressed specifically for different types of perturbations of the system.  相似文献   

15.
16.
Neotyphodium uncinatum and Neotyphodium siegelii are fungal symbionts (endophytes) of meadow fescue (MF; Lolium pratense), which they protect from insects by producing loline alkaloids. High levels of lolines are produced following insect damage or mock herbivory (clipping). Although loline alkaloid levels were greatly elevated in regrowth after clipping, loline-alkaloid biosynthesis (LOL) gene expression in regrowth and basal tissues was similar to unclipped controls. The dramatic increase of lolines in regrowth reflected the much higher concentrations in young (center) versus older (outer) leaf blades, so LOL gene expression was compared in these tissues. In MF-N. siegelii, LOL gene expression was similar in younger and older leaf blades, whereas expression of N. uncinatum LOL genes and some associated biosynthesis genes was higher in younger than older leaf blades. Because lolines are derived from amino acids that are mobilized to new growth, we tested the amino acid levels in center and outer leaf blades. Younger leaf blades of aposymbiotic plants (no endophyte present) had significantly higher levels of asparagine and sometimes glutamine compared to older leaf blades. The amino acid levels were much lower in MF-N. siegelii and MF-N. uncinatum compared to aposymbiotic plants and MF with Epichloë festucae (a closely related symbiont), which lacked lolines. We conclude that loline alkaloid production in young tissue depleted these amino acid pools and was apparently regulated by availability of the amino acid substrates. As a result, lolines maximally protect young host tissues in a fashion similar to endogenous plant metabolites that conform to optimal defense theory.Loline alkaloids (LAs; Hofmeister, 1892; Siegel et al., 1990; TePaske et al., 1993; Blankenship et al., 2001) are protective secondary metabolites produced by some Epichloë and Neotyphodium spp. (epichloae), fungi that live as systemic symbionts in many cool season grasses (Poaceae subfamily Pooideae). The lolines are active against a broad spectrum of insects (Schardl et al., 2007) and are derived from l-Pro (Pro) and l-homoserine (Hse; Blankenship et al., 2005). Mock herbivory (clipping plants) is reported to induce higher levels of lolines in several grass-epichloë symbiota (Craven et al., 2001; Bultman et al., 2004; Gonthier et al., 2008), suggesting that the epichloae have evolved to regulate their metabolism in a manner appropriate for defense of their hosts. However, little is known of the regulation of LA synthesis in symbio and whether these symbionts follow prevailing models for how plants deploy chemical defenses against herbivores (McKey, 1979; Rhoades, 1979; Barto and Cipollini, 2005).The loline-alkaloid biosynthesis (LOL) gene cluster contains nine genes likely to direct LA production (Spiering et al., 2005). Neotyphodium uncinatum contains two highly similar LOL clusters (LOL1 and LOL2), and a single LOL cluster has been found in each of the LA-producing species, Neotyphodium coenophialum, Neotyphodium siegelii, and some strains of Epichloë festucae, among others (Spiering et al., 2005; Kutil et al., 2007). Fermentation cultures of N. uncinatum produce lolines, and studies involving application of labeled precursors and intermediates have almost completely elucidated the LA biosynthetic pathway (Blankenship et al., 2005; Spiering et al., 2005; Faulkner et al., 2006; Schardl et al., 2007). Putative roles of the LOL gene products—based on sequence relationships to known enzyme classes—fit well with the pathway. Furthermore, an RNA interference knockdown of lolC reduces LA levels, and a lolP knockout prevents conversion of N-methylloline to N-formylloline (Spiering et al., 2005, 2008). Expression kinetics of the LOL genes are tightly correlated with each other and with the LA production phase in N. uncinatum cultures (Zhang et al., 2009). This finding raises the question whether and how LOL gene expression in symbio relates to changes in LA levels in response to development and stresses in host plants.LA production in symbio may be influenced by physiological differences among plant tissues and developmental stages, as well as differences in nutritional status and environmental stresses (Kennedy and Bush, 1983; Belesky et al., 1987; Justus et al., 1997; Tong et al., 2006). Given the anti-insect activity of lolines, effects of plant damage on LA levels are of particular interest. Mock herbivory (clipping of leaves) leads to apparent increases in LA concentrations in regrowth tissues of tall fescue (TF; Lolium arundinaceum) symbiotic with N. coenophialum (Bultman et al., 2004; Sullivan et al., 2007) and of meadow fescue (MF; Lolium pratense) symbiotic with N. uncinatum or N. siegelii (Craven et al., 2001). Despite the higher LA levels, however, clipping or damage of TF-N. coenophialum by the herbivore Spodoptera frugiperda (fall armyworm) was reported to elicit only minor, marginally significant (P = 0.052) effects on expression of lolC (Sullivan et al., 2007). A study of the Glyceria striata-Epichloë glyceriae symbiotum demonstrated significantly higher expression of lolC and higher LA production when the grass was artificially damaged, whereas the effect of damage by S. frugiperda on LA concentrations and lolC expression was not significant (Gonthier et al., 2008).Prevailing concepts about how plants deploy chemical defenses include the optimal defense theory (ODT; McKey, 1979; Rhoades, 1979) and the growth differentiation balance hypothesis (GDBH; Barto and Cipollini, 2005). The ODT addresses the distribution of chemical defenses in the plant, predicting that such defenses will be concentrated in tissues that have relatively little means to physically inhibit herbivory (e.g. in young tissues) and are important in the fitness of the plant. The GBDH addresses the location of biosynthesis and predicts that mature tissues are more likely to produce secondary metabolites than are actively growing tissues, which instead need to use resources for biomass production. It is intriguing to consider whether the epichloae obey the predictions of ODT and GDBH, considering that many epichloae protect their hosts by synthesizing insecticidal alkaloids, but they are also evolutionarily derived from plant-pathogenic fungi (Moon et al., 2004) and do not always enhance host fitness (Faeth et al., 2004). In order to address these questions, it is necessary to understand how secondary metabolism of the epichloae is regulated in symbio. The production of lolines in MF-N. uncinatum and MF-N. siegelii is an ideal test case because the lolines accumulate to very high levels—up to 1.9% dry weight—in regrowth of clipped plants (Craven et al., 2001). Here, we test the hypotheses that LOL gene expression and substrate availability correlate with LA levels in younger versus older leaf tissues and in response to clipping in MF-N. uncinatum and MF-N. siegelii symbiota.  相似文献   

17.
Microorganisms can account for up to 60% of the fresh weight of marine sponges. Marine sponges have been hypothesized to serve as accumulation spots of particular microbial communities, but it is unknown to what extent these communities are directed by the organism or the site or occur randomly. To address this question, we assessed the composition of specific bacterial communities associated with Aplysina fulva, one of the prevalent sponge species inhabiting Brazilian waters. Specimens of A. fulva and surrounding seawater were collected in triplicate in shallow water at two sites, Caboclo Island and Tartaruga beach, Búzios, Brazil. Total community DNA was extracted from the samples using “direct” and “indirect” approaches. 16S rRNA-based PCR-denaturing gradient gel electrophoresis (PCR-DGGE) analyses of the total bacterial community and of specific bacterial groups—Pseudomonas and Actinobacteria—revealed that the structure of these assemblages in A. fulva differed drastically from that observed in seawater. The DNA extraction methodology and sampling site were determinative for the composition of actinobacterial communities in A. fulva. However, no such effects could be gleaned from total bacterial and Pseudomonas PCR-DGGE profiles. Bacterial 16S rRNA gene clone libraries constructed from directly and indirectly extracted DNA did not differ significantly with respect to diversity and composition. Altogether, the libraries encompassed 15 bacterial phyla and the candidate division TM7. Clone sequences affiliated with the Cyanobacteria, Chloroflexi, Gamma- and Alphaproteobacteria, Actinobacteria, Bacteroidetes, and Acidobacteria were, in this order, most abundant. The bacterial communities associated with the A. fulva specimens were distinct and differed from those described in studies of sponge-associated microbiota performed with other sponge species.The phylum Porifera (sponges) consists of benthic (sessile) organisms that occur primarily in marine environments at different depths (26). Sponges are classified into three groups, namely, the Calcarea (calcareous sponges), Hexactinellida (glass sponges), and Demospongiae (5, 26). The group Demospongiae, also called demosponges, encompasses 95% of the ca. 5,500 living sponge species described thus far (5). As typical filter feeders, demosponges are the prime bacterial filters of the sea. They are capable of pumping thousands of liters of water per day (23), using prokaryotic microorganisms as the main source of food (1, 43, 47). In addition to demosponges feeding on microorganisms, the presence of bacteria in high density in internal sponge layers (mesohyl) indicates that a selective process favoring particular prokaryotes, involving microbe-sponge interactions, is likely to occur (64). Furthermore, the dawn of the interactions between Prokarya and higher organisms may actually lie in the demosponges, whose origin is estimated to date back to 550 million years ago (5, 33).Putative interactions between demosponges and microorganisms, presumably mostly consisting of Bacteria and Archaea, were first demonstrated by transmission electron microscopy (TEM), where high amounts of microorganisms were observed in the mesohyl (1, 14, 16, 64). Hence, these bacterium-rich sponges have been termed “bacteriosponges” (46). While investigating 11 taxonomically different demosponges using TEM, Vacelet and Donadey (64) identified two different sponge types in respect of their association with bacteria. Sponges with thick mesohyl contained abundant, dense, and morphologically diverse microbial communities (i.e., bacteriosponge), while those with a well-developed aquiferous system and low-density mesohyl contained few bacterial cells and typically only single bacterial morphotypes. The two types have recently been called “high-microbial-abundance” (HMA) and “low-microbial-abundance” (LMA) sponges, respectively (23). In HMA sponges, bacterial densities may reach 108 to 1010 bacterial cells per g (wet weight) of sponge, exceeding seawater concentrations by 2 to 4 orders of magnitude (15, 23). Based on the analysis of 16S rRNA genes, over 15 bacterial phyla have thus far been reported to occur in association with marine sponges (11, 23, 56). Among these are typical sponge-associated bacteria such as members of the Cyanobacteria, Chloroflexi, Proteobacteria, Acidobacteria, Verrucomicrobia, and the candidate phyla “Poribacteria” and TM6 (14, 30, 51, 56, 60, 68).Increasing research interest in the sponge-associated microbiota has emerged in the past few years, mainly due to the in spongium production of an enormous diversity of biologically active secondary metabolites (56). Recent studies suggest that certain bioactive compounds retrieved from marine sponges—such as complex polypeptides and nonribosomal peptides—are likely to be synthesized by the symbiont bacteria (27, 41, 42). Such bioactive secondary metabolites offer great promise for use in biotechnology and medicine (3, 22, 27, 41, 42, 51, 59). In particular, cytotoxic compounds, i.e., antitumoral substances and polyketides, may find application in anticancer therapies (13, 42, 51). Recent investigations revealed the presence of dibromotyrosine-derived metabolites in Aplysina fulva (Pallas, 1766) specimens collected along the Brazilian shore (39). However, a putative role of microbial symbionts in the production of such metabolites, commonly found to display biological activity, remains to be evaluated.Despite the global-scale occurrence of sponges in Earth''s marine ecosystems, the investigation of their associated bacterial communities has thus far been restricted only to certain areas (1, 11, 13, 14, 27, 54, 58, 68). To our knowledge, no studies have been conducted, to date, on sponge-associated microbes in subtropical, South Atlantic open shore waters. In the present study, we assess the diversity and composition of the bacterial community associated with the demosponge A. fulva collected at two different sites at the Brazilian shore. A suite of tools, ranging from plate count estimations and TEM to sponge DNA-based analyses of bacterial 16S rRNA genes, was used. We hypothesized that a distinct bacterial community occurs in A. fulva, which is different from that in the surrounding bulk water, as well as from those in other sponge species.  相似文献   

18.
19.
When plants are infested by herbivores, they emit herbivore-induced plant volatiles (HIPVs) that attract carnivorous natural enemies of herbivores. Furthermore, there are increasing evidences that defenses of intact plants against herbivores are primed when exposed to HIPVs. We previously reported that lima bean leaf volatiles induced by the herbivorous mites Tetranychus urticae primed two T. urtiae-induced indirect defenses in neighboring conspecific plants: HIPV emission and extrafloral nectar (EFN) secretion. An intriguing unanswered question is whether the durations of these two defenses are the same. Here, we show that the durations of the two defenses were the same for up to two days after the initiation of T. urticae damage. The two induced primed defense would act as a battery of defense in exposed plants.Key Words: herbivore-induced plant volatiles, indirect, defense, induced response, plant-plant interaction, primingWhen infested by herbivores, plants defend themselves indirectly by emitting herbivore-induced plant volatiles (HIPVs). One of the ecological functions of HIPVs is to attract carnivorous natural enemies of the herbivores.1,2 Recently, it was reported that the emission of HIPVs primed defenses against herbivores in neighboring intact plants.37 Thus, HIPVs also mediate interactions between infested and intact plants.8 The enhanced defense in response to HIPVs in intact plants is called ‘priming’, which has been studied intensively in plant-pathogen interactions,9 but not so in plant-insect interactions.We previously reported that exposure to HIPVs emitted from lima bean leaves infested by Tetranychus urticae primed HIPV production in detached intact conspecific leaves.3 We also reported that exposure to HIPVs, produced in response to T. urticae damage,4 primed the induced production of extrafloral nectar (EFN; an alternative food source for predators10,11 in lima bean plants. An intriguing question is whether the two primed defenses work as a battery against T. urticae. To answer this, we examined the duration of primed HIPV production by lima bean plants using the same experimental set-up as our previous study of EFN priming by conspecific plants.4For exposure of plants to HIPVs, we used a 60 × 60 × 60 cm cage with two 30 × 30 cm windows on opposite sides of the cage.12 As odor sources, we used eight plants that had been infested with 60 adult T. urticae females per plant for 1 day. Eight uninfested plants were used as control odor sources. Two uninfested plants were placed in a cage with the odor source plants and exposed to either HIPVs or uninfested plant volatiles (UPVs) for 10 days in a climate-controlled room (25 ± 2°C, 60–70% RH, 16:8; L:D).A Y-tube olfactometer13 was used to examine the response of the predators to HIPVs. Adult female P. persimilis were randomly selected from a colony and individually positioned at the beginning of the iron wire. When test mites reached the end of one arm of the olfactometer, their choice was recorded. We tested the olfactory responses of the predator toward (1) plants infested by T. urticae for two days after exposure to UPVs vs. plants infested by T. urticae for two days after exposure to HIPVs, and (2) plants infested by T. urticae for four days after exposure to UPVs vs. plants infested by T. urticae for four days after exposure to HIPVs.HIPV-exposed plants attracted more predators than UPV-exposed plants in a Y-tube olfactometer when infested by T. urticae for two days (Fig. 1A). By contrast, the predators did not distinguish between HIPV- and UPV-exposed plants when infested by T. urticae for four days (Fig. 1B). Our previous study showed that HIPV-exposed plants secreted significantly larger amounts of EFN secretion than UPV-exposed plants infested by T. urticae for two days under the same experimental condition as in this study.4 However, the difference was not significant when they were infested for four days.4Open in a separate windowFigure 1The olfactory response of P. persimilis females to volatiles from the odor-exposed plants, as determined in a Y-tube olfactometer: (A) plants infested by T. urticae for two days after exposure to UPVs (UPV-exposed—T. urticae 2d) vs. plants infested by T. urticae for two days after exposure to HIPVs (HIPV-exposed—T. urticae 2d), and (B) plants infested by T. urticae for four days after exposure to UPVs (UPV-exposed—T. urticae 4d) vs. plants infested by T. urticae for four days after the exposure to HIPVs (HIPV-exposed—T. urticae 4d). Asterisks beside each bar indicate a significant difference between the first trifoliate leaves and the primary leaves. Asterisks beside a bar indicate a significant difference (binomial test: p < 0.001).Lima bean plants increase the amount of endogenous jasmonic acid after exposure to HIPVs.14 Jasmonic acid, an important plant hormone regulating a defense signaling pathway against herbivores and pathogens,15,16 is reported to be involved in the induction of both volatile emission17,18 and EFN secretion19 in response to T. urticae damage in lima bean plants. The increase of endogenous jasmonic acid in HIPV-exposed plants may partly explain the simultaneous priming of the two defenses.In this study, we showed that the durations of priming of two indirect defenses were roughly the same for up to two days. Priming of these two indirect defenses would thus be a battery of defense at the outset of T. urticae damage. Further study is necessarily to test whether the primed battery of induced defense increases the fitness of the exposed plants.  相似文献   

20.
Inhibition of chloroplast reactions with phenylmercuric acetate   总被引:1,自引:1,他引:0       下载免费PDF全文
Phenylmercuric acetate is a selective inhibitor of the photosynthetic activities of isolated spinach (Spinacia oleracea) chloroplasts. At 5 μm concentration of phenylmercuric acetate, photophosphorylation is inhibited. At 33 μm phenylmercuric acetate, ferredoxin is inactivated. Ferredoxin-NADP oxidoreductase is 50% inhibited at 100 μm phenylmercuric acetate. Photosystem II reactions are 50% inhibited at 150 μm phenylmercuric acetate and very much higher cooncentrations—500 μm—are needed to approach complete inhibition. Phenylmercuric acetate inhibition of photosystem II appears to be selective, blocking a site between the 3-(3,4-dichlorophenyl)-1,1-dimethyl urea sensitive site and the site inactivated by high concentrations of tris buffer.  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