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 46 毫秒
1.
2.
NEP1 (necrosis‐ and ethylene‐inducing peptide 1)‐like proteins (NLPs) have been identified in a variety of taxonomically unrelated plant pathogens and share a common characteristic of inducing responses of plant defense and cell death in dicotyledonous plants. Even though some aspects of NLP action have been well characterized, nothing is known about the global range of modifications in proteome and metabolome of NLP‐treated plant cells. Here, using both proteomic and metabolomic approaches we were able to identify the global molecular and biochemical changes in cells of Nicotiana benthamiana elicited by short‐term treatment with MpNEP2, a NLP of Moniliophthora perniciosa, the basidiomycete responsible for the witches' broom disease on cocoa (Theobroma cacao L.). Approximately 100 protein spots were collected from 2‐DE gels in each proteome, with one‐third showing more than twofold differences in the expression values. Fifty‐three such proteins were identified by mass spectrometry (MS)/MS and mapped into specific metabolic pathways and cellular processes. Most MpNEP2 upregulated proteins are involved in nucleotide‐binding function and oxidoreductase activity, whereas the downregulated proteins are mostly involved in glycolysis, response to stress and protein folding. Thirty metabolites were detected by gas spectrometry (GC)/MS and semi‐quantified, of which eleven showed significant differences between the treatments, including proline, alanine, myo‐inositol, ethylene, threonine and hydroxylamine. The global changes described affect the reduction‐oxidation reactions, ATP biosynthesis and key signaling molecules as calcium and hydrogen peroxide. These findings will help creating a broader understanding of NLP‐mediated cell death signaling in plants.  相似文献   

3.
The interaction amongst papain-like cysteine-proteases (PLCP) and their substrates and inhibitors, such as cystatins, can be perceived as part of the molecular battlefield in plant-pathogen interaction. In cacao, four cystatins were identified and characterized by our group. We identified 448 proteases in cacao genome, whereof 134 were cysteine-proteases. We expressed in Escherichia coli a PLCP from cacao, named TcCYSPR04. Immunoblottings with anti-TcCYSPR04 exhibited protein increases during leaf development. Additional isoforms of TcCYSPR04 appeared in senescent leaves and cacao tissues infected by Moniliophthora perniciosa during the transition from the biotrophic to the saprophytic phase. TcCYSPR04 was induced in the apoplastic fluid of Catongo and TSH1188 cacao genotypes, susceptible and resistant to M. perniciosa, respectively, but greater intensity and additional isoforms were observed in TSH1188. The fungal protein MpNEP induced PLCP isoform expression in tobacco leaves, according to the cross reaction with anti-TcCYSPR04. Several protein isoforms were detected at 72 hours after treatment with MpNEP. We captured an active PLCP from cacao tissues, using a recombinant cacao cystatin immobilized in CNBr-Sepharose. Mass spectrometry showed that this protein corresponds to TcCYSPR04. A homology modeling was obtained for both proteins. In order to become active, TcCYSPR04 needs to lose its inhibitory domain. Molecular docking showed the physical-chemical complementarities of the interaction between the cacao enzyme and its inhibitor. We propose that TcCYSPR04 and its interactions with cacao cystatins are involved in the senescence and necrosis events related to witches’ broom symptoms. This molecular interaction may be the target for future interventions to control witches'' broom disease.  相似文献   

4.
5.
Programmed cell death (PCD) plays a key role in plant responses to pathogens, determining the success of infection depending on the pathogen lifestyle and on which participant of the interaction triggers cell death. The hemibiotrophic basidiomycete Moniliophthora perniciosa is the causal agent of witches' broom disease of Theobroma cacao L. (cacao), a serious constraint for production in South America and the Caribbean. It has been hypothesized that M. perniciosa pathogenesis involves PCD, initially as a plant defence mechanism, which is diverted by the fungus to induce necrosis during the dikaryotic phase of the mycelia. Here, we evaluated whether the expression of a cacao anti‐apoptotic gene would affect the incidence and severity of M. perniciosa infection using the ‘Micro‐Tom’ (MT) tomato as a model. The cacao Bax‐inhibitor‐1 (TcBI‐1) gene, encoding a putative basal attenuator of PCD, was constitutively expressed in MT to evaluate function. Transformants expressing TcBI‐1, when treated with tunicamycin, an inducer of endoplasmic reticulum stress, showed a decrease in cell peroxidation. When the same transformants were inoculated with the necrotrophic fungal pathogens Sclerotinia sclerotiorum, Sclerotium rolfsii and Botrytis cinerea, a significant reduction in infection severity was observed, confirming TcBI‐1 function. After inoculation with M. perniciosa, TcBI‐1 transformant lines showed a significant reduction in disease incidence compared with MT. The overexpression of TcBI‐1 appears to affect the ability of germinating spores to penetrate susceptible tissues, restoring part of the non‐host resistance in MT against the S‐biotype of M. perniciosa.  相似文献   

6.
Three cystatin open reading frames named TcCys1, TcCys2 and TcCys3 were identified in cDNA libraries from compatible interactions between Theobroma cacao (cacao) and Moniliophthora perniciosa. In addition, an ORF named TcCys4 was identified in the cDNA library of the incompatible interaction. The cDNAs encoded conceptual proteins with 209, 127, 124, and 205 amino acid residues, with a deduced molecular weight of 24.3, 14.1, 14.3 and 22.8 kDa, respectively. His-tagged recombinant proteins were purified from Escherichia coli expression, and showed inhibitory activities against M. perniciosa. The four recombinant cystatins exhibited K i values against papain in the range of 152–221 nM. Recombinant TcCYS3 and TcCYS4 immobilized in CNBr–Sepharose were efficient to capture M. perniciosa proteases from culture media. Polyclonal antibodies raised against the recombinant TcCYS4 detected that the endogenous protein was more abundant in young cacao tissues, when compared with mature tissues. A ~85 kDa cacao multicystatin induced by M. perniciosa inoculation, MpNEP (necrosis and ethylene-inducing protein) and M. perniciosa culture supernatant infiltration were detected by anti-TcCYS4 antibodies in cacao young tissues. A direct role of the cacao cystatins in the defense against this phytopathogen was proposed, as well as its involvement in the development of symptoms of programmed cell death.  相似文献   

7.
Taxonomy : Moniliophthora roreri (Cif.) H.C. Evans et al. 1978 ; Phylum Basidiomycota; Class Agaricomycetes; Order Agaricales; Family Marasmiaceae; Genus Moniliophthora. Biology : Moniliophthora roreri attacks Theobroma and Herrania species causing frosty pod rot. Theobroma cacao (cacao) is the host of major economic concern. Moniliophthora roreri is a hemibiotroph with a long biotrophic phase (45–90 days). Spore masses, of apparent asexual origin, are produced on the pod surface after initiation of the necrotrophic phase. Spores are spread by wind, rain and human activity. Symptoms of the biotrophic phase can include necrotic flecks and, in some cases, pod malformation, but pods otherwise remain asymptomatic. Relationship to Moniliophthora perniciosa : Moniliophthora roreri and Moniliophthora perniciosa, causal agent of witches’ broom disease of cacao, are closely related. Their genomes are similar, including many of the genes they carry which are considered to be important in the disease process. Moniliophthora perniciosa, also a hemibiotroph, has a typical basidiomycete lifestyle and morphology, forming clamp connections and producing mushrooms. Basidiospores infect meristematic tissues including flower cushions, stem tips and pods. Moniliophthora roreri does not form clamp connections or mushrooms and infects pods only. Both pathogens are limited to the Western Hemisphere and are a threat to cacao production around the world. Agronomic importance : Disease losses caused by frosty pod rot can reach 90% and result in field abandonment. Moniliophthora roreri remains in the invasive phase in the Western Hemisphere, not having reached Brazil, some islands within the Caribbean and a few specific regions within otherwise invaded countries. Disease management : The disease can be managed by a combination of cultural (for example, maintenance of tree height and removal of infected pods) and chemical methods. These methods benefit from regional application, but can be cost prohibitive. Breeding for disease resistance offers the greatest potential for frosty pod rot management and new tolerant materials are becoming available.  相似文献   

8.
The basidiomycete Ustilago maydis, the causative agent of corn smut disease, has emerged as a model organism for dimorphism and fungal phytopathogenicity. In this work, we line out the key conserved enzymes for β-carotene biosynthesis encoded by the U. maydis genome and show that this biotrophic fungus accumulates β-carotene. The amount of this pigment depended on culture pH and aeration but was not affected by light and was not increased by oxidative stress. Moreover, we identified the U. maydis gene, cco1, encoding a putative β-carotene cleavage oxygenase. Heterologous overexpression and in vitro analyses of purified enzyme demonstrated that Cco1 catalyzes the symmetrical cleavage of β-carotene to yield two molecules of retinal. Analyses of β-carotene and retinal contents in U. maydis cco1 deletion and over-expression strains confirmed the enzymatic function of Cco1, and revealed that Cco1 determines the β-carotene content. Our data indicate that carotenoid biosynthesis in U. maydis is carried out to provide retinal rather than to deliver protective pigments. The U. maydis genome also encodes three potential opsins, a family of photoactive proteins that use retinal as chromophore. Two opsin genes showed different light-regulated expression patterns, suggesting specialized roles in photobiology, while no mRNA was detected for the third opsin gene in the same experiments. However, deletion of the cco1 gene, which should abolish function of all the retinal-dependent opsins, did not affect growth, morphology or pathogenicity, suggesting that retinal and opsin proteins play no relevant role in U. maydis under the tested conditions.  相似文献   

9.
Oxalic acid has been shown as a virulence factor for some phytopathogenic fungi, removing calcium from pectin and favoring plant cell wall degradation. Recently, it was published that calcium oxalate accumulates in infected cacao tissues during the progression of Witches’ Broom disease (WBD). In the present work we report that the hemibiotrophic basidiomycete Moniliophthora perniciosa, the causal agent of WBD, produces calcium oxalate crystals. These crystals were initially observed by polarized light microscopy of hyphae growing on a glass slide, apparently being secreted from the cells. The analysis was refined by Scanning electron microscopy and the compositon of the crystals was confirmed by energy-dispersive x-ray spectrometry. The production of oxalate by M. perniciosa was reinforced by the identification of a putative gene coding for oxaloacetate acetylhydrolase, which catalyzes the hydrolysis of oxaloacetate to oxalate and acetate. This gene was shown to be expressed in the biotrophic-like mycelia, which in planta occupy the intercellular middle-lamella space, a region filled with pectin. Taken together, our results suggest that oxalate production by M. perniciosa may play a role in the WBD pathogenesis mechanism.  相似文献   

10.
Streptomyces albovinaceus, Streptomyces caviscabies, Streptomyces griseus, Streptomyces setonii, and Streptomyces virginiae selected as antagonists of Moniliophthora (ex Crinipellis) perniciosa, the causal agent of cacao Witches’ broom, were examined in vitro to detect production of chitinases, β-1,3-glucanases, and cellulases. All the species produced chitinases, but not β-1,3-glucanases or cellulases, when grown on a liquid mineral medium containing glucose, colloidal chitin, or cell walls of M. perniciosa as a carbon source. There were no quantitative differences among species in the production of chitinase, however, the germination inhibition of basidiospores of M. perniciosa was higher when they were cultivated using glucose as a carbon source, followed by colloidal chitin and cell walls. All the species also produced hydroxymate type siderophores in similar quantities, and the quantity of siderophores did not correlate with the inhibition of basidiospore germination. The germination inhibition was more pronounced when S. albovinaceus, S. griseus, and S. virginiae were cultivated on iron-deficient medium, suggesting involvement of siderophores in the antagonism by these species of actinomycetes.  相似文献   

11.
As a notable biocontrol agent, Trichoderma harzianum can antagonize a diverse array of phytopathogenic fungi, including Botrytis cinerea, Rhizoctonia solani and Fusarium oxysporum. Elucidating the biocontrol mechanism of T. harzianum in response to the pathogens enables it to be exploited in the control of plant diseases. Two-dimensional gel electrophoresis (2-DE) was performed to obtain secreted protein patterns of T. harzianum ETS 323, grown in media that contained glucose, a mixture of glucose and deactivated B. cinerea mycelia, deactivated B. cinerea mycelia or deactivated T. harzianum mycelia. Selected protein spots were identified using liquid chromatography–tandem mass spectrometry (LC–MS/MS). Ninety one out of 100 excised protein spots were analyzed and some proteins were sequence identified. Of these, one l-amino acid oxidase (LAAO) and two endochitinases were uniquely induced in the media that contained deactivated B. cinerea mycelia as the sole carbon source. Activities of the cell wall-degrading enzymes (CWDEs), including β-1,3-glucanases, β-1,6-glucanases, chitinases, proteases and xylanases, were significantly higher in media with deactivated B. cinerea mycelia than in other media. This finding suggests that the cell wall of B. cinerea is indeed the primary target of T. harzianum ETS 323 in the biocontrol mechanism. The possible roles of LAAO and xylanase were also discussed.  相似文献   

12.
13.
Witches’ broom disease (WBD) of cacao, caused by the hemibiotrophic fungus, Crinipellis perniciosa, exhibits a succession of symptoms that are caused by the biotrophic phase of the fungus. However, the study of this biotrophic phase is limited by its exclusive growth inside the plant or in the presence of callus. Here we report for the first time a method for the growth and maintenance of the biotrophic-like phase of C. perniciosa on a defined medium with metabolites found in the diseased tissues. Our results suggest that glycerol is a key carbon source for this interaction. This is a crucial achievement toward understanding the biology of this fungus during the infectious phase of WBD.  相似文献   

14.
Isolation and identification of substances having an activity to stimulate the fruiting body formation of Schizophyllum commune were attempted. The active principles in its mycelia were divided into four fractions by sequential purification with silica gel column and reverse-phase HPLC column chromatography. By infrared spectra and thin-layer chromatography, the active substances in these four fractions were revealed as cerebrosides. About 0.1 μg of the cerebroside fractions showed a discriminative stimulating activity on S. commune when tested by the method these authors adopted. The active substance in the fraction II was N-2′-hydroxypalmitoyl-1-O-glucosyl-nonadecasphingadienine. The cerebrosides from pea seeds and Fusicoccum amygdali showed the similar activity on S. commune, but some commercial synthetic cerebrosides and cerebrosides from bovine and porcine brains exhibited no stimulating activity. Only definite cerebrosides with special structures seem to be able to induce the fruiting of S. commune.  相似文献   

15.
Nep1-like proteins (NLPs) are a novel family of microbial elicitors of plant necrosis that induce a hypersensitive-like response in dicot plants. The spatial structure and role of these proteins are yet unknown. In a paper published in BMC Plant Biology (2008; 8:50) we have proposed that the core region of Nep1-like proteins (NLPs) belong to the Cupin superfamily. Based on what is known about the Cupin superfamily, in this addendum to the paper we discuss how NLPs could form oligomers.Key words: quaternary structure, necrosis and ethylene inducing proteins, NLPs, MpNEP1, MpNEP2, NPP1, Moniliophthora perniciosa, Phytophthora parasiticaCupins may be organized as monomers, dimers, hexamers and octamers of β-barrel domains.1 To the best of our knowledge trimers have not been detected yet. The interaction of two monomers building up a dimeric structure is basically performed by three types of interactions: hydrophobic interactions between β-strands in different subunits, salt bridges and hydrogen bonds between β-strands. In cupin dimers, the hydrophobic interactions occur between two βI strands in different subunits (Fig. 1A and B). This strand represents the central axis of rotation of the dimer as one residue in βI interacts with the corresponding residue in the other subunit (Fig. 1B). Therefore, all residues in βI must be hydrophobic, as one residue interacts with the other subunit and the next one in the sequence interacts with the interior of the protein. Charged residues in βI would disrupt such interactions. Most cupin dimers have strong hydrophobic residues such as tryptophan (W), phenylalanine (F) and methionine (M) pointing towards the own subunit (↓), while small hydrophobic residues such as leucine (L), isoleucine (I), and valine (V) point to the other subunit (↑). A particular case is leucine that interacts with other subunits, for instance, βI = liaW (positions 217–220 in Fig. 1B) and βI = LVsw of type I and II NLP consensuses, respectively. Therefore, the pattern of hydropathicity suggests that the side chain orientation is βI = l217 ↑ i218 ↓ a219 ↑ W220 ↓ d221 ↑. However we observe that just after βI there is a charged residue (aspartate D221) which would point outwards disrupting the dimer or at least making it less stable. It is interesting to observe that the requirement for a negatively charged residue at this last position is very high: 96% of all type I NLPs contains an aspartate (D) or glutamate (E) indicating an important role for it, maybe in avoiding dimerization of the NLPs. A second interesting hypothesis is as follows: several cupins are oxygenases, decarboxylases, etc. and use a negatively charged residue, such as aspartate or glutamate as proton donor.1 Now, if the alternate pattern of side chains of the residues is βI = l217 ↓ i218 ↑ a219 ↓ W220 ↑ d221 ↓, instead of the previous one, then the aspartate or glutamate residue would point to the hydrophobic pocket and would be positioned to interact with the metal ion, as in cupins with enzymatic activity. However, there are no experimental evidences that the NLPs have enzymatic activity.Open in a separate windowFigure 1(A) Three-dimensional structure prediction for type I NLP consensus, (B) Interface between two βI strands in type I NLP consensus. From the left to the right: EF-coil with the conserved residue H162, βC and βH strands (superposed) with the conserved histidines H133 and H135 in βC, H193 and leucine L195 in βH, W220 in βI and W118 in βB. The strands in the right subunit follow the same pattern but rotated.The second type of interaction is salt bridges between charged residues in different subunits. Analyzing all interacting side chains in the 1VJ2 protein (dimer), we verify that the charged side chains of N35 and E57 (numbers in original structure) are only 2.72 Å apart. In the NLPs, this corresponds to N10836% (Q10860%) at the border of βB and E13898%. The negatively charged residue D125 helps to correct the orientation of the subunits in relation to each other avoiding any disorientation. The high conservation level of these residues suggests that NLPs are dimeric structures. However, as we will see next, only hydrophobic and charged interactions are not enough to build a dimer.Garcia et al. (2007)2 have used small angle X-ray scattering (SAXS) to show that, in solution, at low concentrations (<2 mg/ml) the two copies of the NLPs of Moniliophthora perniciosa, MpNEP1 and MpNEP2, exist as dimers and monomers, respectively. The same technique showed that at higher concentrations, >5 mg/ml, both proteins exist as dimers, as is the case for PpNPP1.2 They also reported, based on electrophoresis analysis, that PpNPP1 and MpNEP1 exist as oligomers and MpNEP2 as monomers.2 However, experiments with the PpNPP1 in size exclusion chromatography using myoglobin as size standard suggest that PpNPP1 is a monomer.3 Figure 2 compares MpNEP1, MpNEP2 and PpNPP1, where the most relevant differences in sequence are marked with asterisks (*) and are possibly related to the differences in oligomeric properties between MpNEP1 and PpNPP1 with MpNEP2. These positions are methionine M27 and leucine L35, which occur only in MpNEP2, glycine G250, which occurs only in MpNEP2 and NEP1 (Fusarium oxysporum) and lysine K31, which occurs only MpNEP2, BAB04114 (Bacillus halodurans) and AAU23136 (Bacillus licheniformis). The other residues are aspartate D28, which occurs 9 times and alanine A37 which occurs 7 times of all investigated NLPs. Thus, the sequence mdHDkiakl at the start of the NLPs seems to explain the monomeric state of MpNEP2, although at higher concentrations they form dimers. Besides the weak hydrophobic interactions, dimeric cupins and bicupins (two β barrels in the same sequence building up a dimeric-like 4d-structure) are stable structures (see Fig. 1 in ref. 4). By aggregating the first β-strand in the start domain of one β-barrel to the ABIDG β-sheet of the other β-barrel, composing a big ABIDGY β-sheet (Y is the first β-strand). For instance, using the bicupin 1L3J (oxalate decarboxylase) as template, the low confidence level β-strand at position 26–33 (v in H29D30 avv) in type I NLPs corresponds to the first β-strand. Since this proceeds from both barrels they can build a stable structure (see Fig. 1 in ref. 4). The quaternary structure is related to the presence of interaction residues in the BID β-sheet of the cupin structure. These are present in the NLPs and would enable them to form dimers.Open in a separate windowFigure 2Alignment of type I NLP consensus, PpNPP1, MpNEP1 and MpNEP2. Solid line boxes are β-strands, double line boxes are α-helices. The sequence positions marked with asterisks (*) are possibly related to the differences in oligomeric properties between MpNEP1 and PpNPP1 with MpNEP2.  相似文献   

16.
17.
Mitochondrial intermediate peptidase (MIP) is a component of the mitochondrial protein import machinery required for maturation of nuclear-encoded precursor proteins targeted to the mitochondrial matrix or inner membrane. We previously characterized this enzyme in rat (RMIP) and Saccharomyces cerevisiae (YMIP) and showed that MIP activity is essential for mitochondrial function in yeast. We have now defined the structure of a new MIP homologue (SMIP) from the basidiomycete fungus Schizophyllum commune. SMIP includes 4 exons of 523, 486, 660, and 629 bp separated by 3 short introns. The predicted SMIP, YMIP, and RMIP sequences share 31-37% identity and 54-57% similarity over 700 amino acids. When SMIP and RMIP were expressed in a yeast mip1Δ mutant, they were both able to rescue the respiratory-deficient phenotype caused by genetic inactivation of YMIP, indicating that the function of this enzyme is conserved in eukaryotes. Moreover, the MIP sequences show 20-24% identity and 40-47% similarity to a family of oligopeptidases from bacteria, yeast, and mammals. MIP and these proteins are characterized by a highly conserved motif, F-H-E-X-G-H-(X)2-H-(X)12-G-(X)5-D-(X)2-E-X-P-S-(X)3-E-X, centered around a zinc-binding site and appear to represent a new family of genes associated with proteolytic processing in the mitochondrial and cytosolic compartments.  相似文献   

18.
19.
20.
The nucleotide (nt) sequences of the Sc3 and Sc4 genes of the filamentous fungus Schizophyllum commune, and the deduced amino acid (aa) sequences, were determined; moreover, the previously published sequence for the ScI gene [Dons et al., EMBO J. 3 (1984) 2101–2106] was corrected. All three independently isolated genes were found to have similar structures and nt sequences of their coding regions. At the aa level the homology is 43–62% (63–69% in the C-terminal parts of the proteins), the hydrophobic aa predominate and the hydrophobicity patterns are similar. All three proteins contain leader sequences and eight cysteines among about 110 aa, conserved at the same positions. Yet these genes are differentially regulated: Sc1 and Sc4 are only expressed at high levels in fruiting dikaryons, whereas Sc3 is highly expressed in both monokaryons and dikaryons, independent from fruiting.  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