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.
3.
Whole-cell matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (whole-cell MALDI-TOF MS) has been widely adopted as a useful technology in the identification and typing of microorganisms. This study employed the whole-cell MALDI-TOF MS to identify and differentiate wild-type and mutants containing constructed single gene mutations of Burkholderia pseudomallei, a pathogenic bacterium causing melioidosis disease in both humans and animals. Candidate biomarkers for the B. pseudomallei mutants, including rpoS, ppk, and bpsI isolates, were determined. Taxon-specific and clinical isolate-specific biomarkers of B. pseudomallei were consistently found and conserved across all average mass spectra. Cluster analysis of MALDI spectra of all isolates exhibited separate distribution. A total of twelve potential mass peaks discriminating between wild-type and mutant isolates were identified using ClinProTools analysis. Two peaks (m/z 2721 and 2748 Da) were specific for the rpoS isolate, three (m/z 3150, 3378, and 7994 Da) for ppk, and seven (m/z 3420, 3520, 3587, 3688, 4623, 4708, and 5450 Da) for bpsI. Our findings demonstrated that the rapid, accurate, and reproducible mass profiling technology could have new implications in laboratory-based rapid differentiation of extensive libraries of genetically altered bacteria.  相似文献   

4.
分子马达不等间距四态跃迁模型   总被引:4,自引:0,他引:4  
用主方程方法研究分子马达一维周期性四态不等间距随机跃迁模型,得出稳态情况下分子马达的漂移速度V、扩散系数D及随机参数r,将三者随ATP的浓度[ATP]及外力F的变化进行了曲线拟合,并同大量实验结果进行了比较,定性半定量地分析了分子马达在拖动负载运动时的动力学行为.  相似文献   

5.
A wild-type nucleopolyhedrovirus (NPV) isolate from Spodoptera exigua from Florida (Se-US2) is a variant of the SeMNPV type strain since it has a unique DNA profile but is closely related to other known geographical isolates of SeMNPV. It consists of several genotypic variants, of which seven were identified in a Se-US2 virus stock by a modification of the in vivo cloning method developed by Smith and Crook (Virology 166:240–244, 1988). The US2A variant was the most prevalent genotype, and it was designated the prototype Se-US2 variant, while four of the variants (US2B, US2D, US2F, and US2H) were found at low frequency. US2C and US2E were also very abundant, and their diagnostic bands were easily observed in wild-type isolate restriction endonuclease patterns. The analysis of each variant, compared to the prototype US2A, showed that US2B and US2H presented minor differences, while US2D and US2F contained slightly larger insertions or deletions. Variants US2C and US2E contained major deletions of 21.1 and 14 kb, respectively, mapping at the same genomic region (between 14.5 and 30.2 map units [m.u.] and between 12.8 and 23 m.u., respectively). This is the first report of such deletion mutants in a natural baculovirus population. Variants US2A, US2B, US2D, US2F, and US2H were isolated as pure genotypes, but we failed to clone US2C and US2E in vivo. When these two variants appeared without apparent contamination with any other variant, they lost their pathogenicity for Spodoptera exigua larvae. A further biological characterization showed evidence that these two naturally occurring deletion mutants act as parasitic genotypes in the virus population. Bioassay data also demonstrated that pure US2A is significantly more pathogenic against second-instar S. exigua larvae than the wild-type isolate. The need for precise genotypic characterization of a baculovirus prior to its development as a bioinsecticide is discussed.  相似文献   

6.
The enzyme Erwinia chrysanthemi l-asparaginase (ErA) is an important biopharmaceutical product used in the treatment of acute lymphoblastic leukaemia. Like all proteins, certain asparagine (Asn) residues of ErA are susceptible to deamidation to aspartic acid (Asp), which may be a concern with respect to enzyme activity and potentially to pharmaceutical efficacy. Recombinant ErA mutants containing Asn to Asp changes were expressed, purified and characterised. Two mutants with single deamidation sites (N41D and N281D) were found to have approximately the same specific activity (1,062 and 924 U/mg, respectively) as the wild-type (908 U/mg). However, a double mutant (N41D N281D) had an increased specific activity (1261 U/mg). The N41D mutation conferred a slight increase in the catalytic constant (k cat 657 s?1) when compared to the WT (k cat 565 s?1), which was further increased in the double mutant, with a k cat of 798 s?1. Structural analyses showed that the slight changes caused by point mutation of Asn41 to Asp may have reduced the number of hydrogen bonds in this α-helical part of the protein structure, resulting in subtle changes in enzyme turnover, both structurally and catalytically. The increased α-helical content observed with the N41D mutation by circular dichroism spectroscopy correlates with the difference in k cat, but not K m. The N281D mutation resulted in a lower glutaminase activity compared with WT and the N41D mutant, however the N281D mutation also imparted less stability to the enzyme at elevated temperatures. Taken as a whole, these data suggest that ErA deamidation at the Asn41 and Asn281 sites does not affect enzyme activity and should not be a concern during processing, storage or clinical use. The production of recombinant deamidated variants has proven an effective and powerful means of studying the effect of these changes and may be a useful strategy for other biopharmaceutical products.  相似文献   

7.
8.
The mechanisms of action of phosphate solubilization were studied in the wild-type strain Aspergillus tubingensis and the phenotypic mutants derived from it. The P solubilization activities of these isolates were measured in liquid media using different carbon and nitrogen sources. All the mutants showed higher P solubilization compared to the wild type. Glucose and sucrose significantly promoted P solubilization compared to fructose, lactose, galactose, and xylose. Potassium nitrate significantly increased P solubilization compared to other nitrogen sources such as ammonium sulfate, ammonium nitrate, aspargine, and tryptophan. The P solubilization activity was strongly associated with the production of organic acids, especially succinic acid and acetic acid. The enzyme activities such as acid phosphatase and phytase also increased significantly in mutants compared to the wild type. These results suggested the role of these enzymes in P solubilization apart from the organic acid exudation and H+ pump in A. tubingensis.  相似文献   

9.
The fungus Fusarium fujikuroi (Gibberella fujikuroi MP-C) produces metabolites of biotechnological interest, such as gibberellins, bikaverins, and carotenoids. Gibberellin and bikaverin productions are induced upon nitrogen exhaustion, while carotenoid accumulation is stimulated by light. We evaluated the effect of nitrogen availability on carotenogenesis in comparison with bikaverin and gibberellin production in the wild type and in carotenoid-overproducing mutants (carS). Nitrogen starvation increased carotenoid accumulation in all strains tested. In carS strains, gibberellin and bikaverin biosynthesis patterns differed from those of the wild type and paralleled the expression of key genes for both pathways, coding for geranylgeranyl pyrophosphate (GGPP) and kaurene synthases for the former and a polyketide synthase for the latter. These results suggest regulatory connections between carotenoid biosynthesis and nitrogen-controlled biosynthetic pathways in this fungus. Expression of gene ggs1, which encodes a second GGPP synthase, was also derepressed in the carS mutants, suggesting the participation of Ggs1 in carotenoid biosynthesis. The carS mutations did not affect genes for earlier steps of the terpenoid pathway, such as fppS or hmgR. Light induced carotenoid biosynthesis in the wild type and carRA and carB levels in the wild-type and carS strains irrespective of nitrogen availability.  相似文献   

10.
Hepatitis B virus immune escape mutants have been associated with vaccine failure and reinfection of grafted liver despite immune prophylaxis, but their biological properties remain largely unknown. Transfection of 20 such mutants in a human hepatoma cell line identified many with severe impairment in virion secretion, which can be rescued to various extents by coexpression of wild-type envelope proteins or introduction of a novel glycosylation site. Consistent with their role in maintaining intra- or intermolecular disulfide bonds, cysteine residues within the “a” determinant are critical for virion secretion.  相似文献   

11.
Betaine aldehyde dehydrogenase 2 (BADH2) is believed to be involved in the accumulation of 2-acetyl-1-pyrroline (2AP), one of the major aromatic compounds in fragrant rice. The enzyme can oxidize ω-aminoaldehydes to the corresponding ω-amino acids. This study was carried out to investigate the function of wild-type BADHs and four BADH2 mutants: BADH2_Y420, containing a Y420 insertion similar to BADH2.8 in Myanmar fragrance rice, BADH2_C294A, BADH2_E260A and BADH2_N162A, consisting of a single catalytic-residue mutation. Our results showed that the BADH2_Y420 mutant exhibited less catalytic efficiency towards γ-aminobutyraldehyde but greater efficiency towards betaine aldehyde than wild-type. We hypothesized that this point mutation may account for the accumulation of γ-aminobutyraldehyde/Δ1-pyrroline prior to conversion to 2AP, generating fragrance in Myanmar rice. In addition, the three catalytic-residue mutants confirmed that residues C294, E260 and N162 were involved in the catalytic activity of BADH2 similar to those of other BADHs.  相似文献   

12.
The effects of Newcastle disease virus (NDV) fusion (F) glycoprotein cleavage mutants on the cleavage and syncytium-forming activity of the wild-type F protein were examined. F protein cleavage mutants were made by altering amino acids in the furin recognition region (amino acids 112 to 116) in the F protein of a virulent strain of NDV. Four mutants were made: Q114P replaced the glutamine residue with proline; K115G replaced lysine with glycine; double mutant K115G, R113G replaced both a lysine and an arginine with glycine residues; and a triple mutant, R112G, K115G, F117L, replaced three amino acids to mimic the sequence found in avirulent strains of NDV. All mutants except Q114P were cleavage negative and fusion negative. However, addition of exogenous trypsin cleaved all mutant F proteins and activated fusion. As expected for an oligomeric protein, the fusion-negative mutants had a dominant negative phenotype: cotransfection of wild-type and mutant F protein cDNAs resulted in an inhibition of syncytium formation. The presence of the mutant F protein did not inhibit cleavage of the wild-type protein. Furthermore, evidence is presented that suggests that the mutant protein and the wild-type protein formed heterooligomers. By measuring the syncytium-forming activity of the wild-type protein at various ratios of expression of mutant and wild-type protein, results were obtained that are most consistent with the notion that the size of the functionally active NDV F protein in these assays is a single oligomer, likely a trimer. That a larger oligomer, containing a mix of both wild-type and mutant F proteins, has partial activity cannot, however, be ruled out.  相似文献   

13.
Conditions for superiority of the minimum dispersion estimator over another with respect to the covariance matrix are derived when the vector parameter of a regression model is subject to competing stochastic restrictions. The restrictions may also consist both of a deterministic part and a stochastic part.  相似文献   

14.
IKs channels are voltage dependent and K+ selective. They influence cardiac action potential duration through their contribution to myocyte repolarization. Assembled from minK and KvLQT1 subunits, IKs channels are notable for a heteromeric ion conduction pathway in which both subunit types contribute to pore formation. This study was undertaken to assess the effects of minK on pore function. We first characterized the properties of wild-type human IKs channels and channels formed only of KvLQT1 subunits. Channels were expressed in Xenopus laevis oocytes or Chinese hamster ovary cells and currents recorded in excised membrane patches or whole-cell mode. Unitary conductance estimates were dependent on bandwidth due to rapid channel “flicker.” At 25 kHz in symmetrical 100-mM KCl, the single-channel conductance of IKs channels was ∼16 pS (corresponding to ∼0.8 pA at 50 mV) as judged by noise-variance analysis; this was fourfold greater than the estimated conductance of homomeric KvLQT1 channels. Mutant IKs channels formed with D76N and S74L minK subunits are associated with long QT syndrome. When compared with wild type, mutant channels showed lower unitary currents and diminished open probabilities with only minor changes in ion permeabilities. Apparently, the mutations altered single-channel currents at a site in the pore distinct from the ion selectivity apparatus. Patients carrying these mutant minK genes are expected to manifest decreased K+ flux through IKs channels due to lowered single-channel conductance and altered gating.  相似文献   

15.
16.
In an assay of carbonic anhydrase (CA), NAH14CO3 soltution at the bottom of a sealed vessel releases 14CO2, which diffuses to the top of the vessel to be assimilated by photosynthesizing Chlamydomonas reinhardtii cells that have been adapted to a low-CO2 environment. The assay is initiated by illuminating the cells and is stopped by turning the light off and killing the cells with acid. Enzyme activity was estimated from acid-stable radioactivity. With bovine CA, 1.5 Wilbur-Anderson units (WAU) was consistently measured at 5- to 6-fold above background. Sonicated whole cells of air-adapted wild-type C. reinhardtii had 740 [plus or minus] 12.4 WAU/mg chlorophyll (Chl). Sonicated chloroplasts from a mixotrophically grown wall-less strain, cw-15, had 35.5 [plus or minus] 2.6 WAU/mg Chl, whereas chloroplasts from wall-less external CA mutant strain cia5/cw-15 had 33.8 [plus or minus] 1.9 WAU/mg Chl. Sonicated chloroplasts from the wall-less mutant strain cia-3/cw-15, believed to lack an internal CA, had 2.8 [plus or minus] 3.2 WAU/mg Chl. Sonicated whole cells from cia3/cw-15 had 2.8 [plus or minus] 7.8 WAU/mg Chl. Acetazolamide, ethoxyzolamide, and p-aminomethylbenzene sulfonamide (Mafenide) at 100 [mu]M inhibited CA in sonicated chloroplasts from cia-5/cw-15. Treatment at 80[deg]C for 10 min inhibited this CA activity by 90.8 [plus or minus] 3.6%. Thus, a sensitive 14C assay has confirmed the presence of a CA in cw-15 and cia-5/cw-15 chloroplasts and the lack of a CA in cia-3/cw-15 chloroplasts. Our results indicate that HCO3- is the inorganic carbon species that is accumulated by chloroplasts of Chlamydomonas and that chloroplastic CA is responsible for the majority of internal CA activity.  相似文献   

17.
The minimum dispersion linear unbiased estimators of the vector of parameters in a linear regression model are compared when the parameters of the model are subject to stochastic linear restrictions with different dispersion matrices of the disturbances involved in them.  相似文献   

18.
Escherichia coli has only a single copy of a gene for tRNA6Leu (Y. Komine et al., J. Mol. Biol. 212:579–598, 1990). The anticodon of this tRNA is CAA (the wobble position C is modified to O2-methylcytidine), and it recognizes the codon UUG. Since UUG is also recognized by tRNA4Leu, which has UAA (the wobble position U is modified to 5-carboxymethylaminomethyl-O2-methyluridine) as its anticodon, tRNA6Leu is not essential for protein synthesis. The BT63 strain has a mutation in the anticodon of tRNA6Leu with a change from CAA to CUA, which results in the amber suppressor activity of this strain (supP, Su+6). We isolated 18 temperature-sensitive (ts) mutants of the BT63 strain whose temperature sensitivity was complemented by introduction of the wild-type gene for tRNA6Leu. These tRNA6Leu-requiring mutants were classified into two groups. The 10 group I mutants had a mutation in the miaA gene, whose product is involved in a modification of tRNAs that stabilizes codon-anticodon interactions. Overexpression of the gene for tRNA4Leu restored the growth of group I mutants at 42°C. Replacement of the CUG codon with UUG reduced the efficiency of translation in group I mutants. These results suggest that unmodified tRNA4Leu poorly recognizes the UUG codon at 42°C and that the wild-type tRNA6Leu is required for translation in order to maintain cell viability. The mutations in the six group II mutants were complemented by introduction of the gidA gene, which may be involved in cell division. The reduced efficiency of translation caused by replacement of the CUG codon with UUG was also observed in group II mutants. The mechanism of requirement for tRNA6Leu remains to be investigated.In the universal genetic code, 61 sense codons correspond to 20 amino acids, and the various tRNA species mediate the flow of information from the genetic code to amino acid sequences. Since codon-anticodon interactions permit wobble pairing at the third position, 32 tRNAs, including tRNAfMet, should theoretically be sufficient for a complete translation system. Although some organisms have fewer tRNAs (1), most have abundant tRNA species and multiple copies of major tRNAs. For example, Escherichia coli has 86 genes for tRNA (79 genes identified in reference 14, 6 new ones reported in reference 3, and one fMet tRNA at positions 2945406 to 2945482) that encode 46 different amino acid acceptor species. Although abundant genes for tRNAs are probably required for efficient translation, the significance of the apparently nonessential tRNAs has not been examined.E. coli has five isoaccepting species of tRNALeu. According to the wobble rule, tRNA1Leu recognizes only the CUG codon. The CUG codon is also recognized by tRNA3Leu (tRNA2Leu) and thus tRNA1Leu may not be essential for protein synthesis. Similarly, tRNA6Leu is supposed to recognize only the UUG codon, but tRNA4Leu can recognize both UUA and UUG codons. Thus, tRNA6Leu appears to be dispensable. The existence of an amber suppressor mutation of tRNA6Leu (supP, Su+6) supports this possibility. tRNA6Leu is encoded by a single-copy gene, leuX (supP), and Su+6 has a mutation in the anticodon, which suggests loss of the ability to recognize UUG (26). Why are so many species of tRNALeu required? Holmes et al. (12) examined the utilization of the isoaccepting species of tRNALeu in protein synthesis and showed that utilization differs depending on the growth medium; in minimal medium, isoacceptors tRNA2Leu (cited as tRNA3Leu; see Materials and Methods) and tRNA4Leu are the predominant species that are found bound to ribosomes, but an increased relative level of tRNA1Leu is found bound to ribosomes in rich medium. The existence of tRNA6Leu is puzzling. This isoaccepting tRNA accounts for approximately 10% of the tRNALeu in total-cell extracts. However, little if any tRNA6Leu is found on ribosomes in vivo, and it is also only weakly active in protein synthesis in vitro with mRNA from E. coli (12). It thus appears that tRNA6Leu is only minimally involved in protein synthesis in E. coli.To investigate the role of tRNA6Leu in E. coli, we attempted to isolate tRNA6Leu-requiring mutants from an Su+6 strain. These mutants required wild-type tRNA6Leu for survival at a nonpermissive temperature. We report here the isolation and the characterization of these mutants.  相似文献   

19.
20.
Cuticle function is closely related to the structure of the cutin polymer. However, the structure and formation of this hydrophobic polyester of glycerol and hydroxy/epoxy fatty acids has not been fully resolved. An apoplastic GDSL-lipase known as CUTIN SYNTHASE1 (CUS1) is required for cutin deposition in tomato (Solanum lycopersicum) fruit exocarp. In vitro, CUS1 catalyzes the self-transesterification of 2-monoacylglycerol of 9(10),16-dihydroxyhexadecanoic acid, the major tomato cutin monomer. This reaction releases glycerol and leads to the formation of oligomers with the secondary hydroxyl group remaining nonesterified. To check this mechanism in planta, a benzyl etherification of nonesterified hydroxyl groups of glycerol and hydroxy fatty acids was performed within cutin. Remarkably, in addition to a significant decrease in cutin deposition, mid-chain hydroxyl esterification of the dihydroxyhexadecanoic acid was affected in tomato RNA interference and ethyl methanesulfonate-cus1 mutants. Furthermore, in these mutants, the esterification of both sn-1,3 and sn-2 positions of glycerol was impacted, and their cutin contained a higher molar glycerol-to-dihydroxyhexadecanoic acid ratio. Therefore, in planta, CUS1 can catalyze the esterification of both primary and secondary alcohol groups of cutin monomers, and another enzymatic or nonenzymatic mechanism of polymerization may coexist with CUS1-catalyzed polymerization. This mechanism is poorly efficient with secondary alcohol groups and produces polyesters with lower molecular size. Confocal Raman imaging of benzyl etherified cutins showed that the polymerization is heterogenous at the fruit surface. Finally, by comparing tomato mutants either affected or not in cutin polymerization, we concluded that the level of cutin cross-linking had no significant impact on water permeance.Cuticles are ubiquitous hydrophobic barriers at the surfaces of aerial plant organs. These complex hydrophobic assemblies consist of a biopolymer, cutin, coated and filled with waxes and can also comprise embedded cell wall polysaccharides. Waxes comprise solvent-soluble aliphatic molecules with long hydrocarbon chains, terpenes, and steroids (Kunst and Samuels, 2003; Nawrath, 2006; Pollard et al., 2008; Samuels et al., 2008; Schreiber, 2010; Lee and Suh, 2015). Cutin is an insoluble polyester of ω- and mid-chain hydroxy C16 and C18 fatty acids. Glycerol has also been described as a ubiquitous cutin monomer (Graça et al., 2002; Pollard et al., 2008). In some cuticles, a hydrophobic polymer that is resistant to alkaline hydrolysis (i.e. cutan) has been observed (Gupta et al., 2006; Li-Beisson et al., 2010).Cutin fulfills multiple functions in plants, such as the control of nonstomatal water loss (Sieber et al., 2000) and the permeation of gases and solutes (Kersteins, 1996; Schreiber, 2010). Cutin also plays an essential role in the regulation of cell adhesion during plant development by preventing organ fusion, as observed in Arabidopsis (Arabidopsis thaliana) mutants with cuticle defects (Sieber et al., 2000; Nawrath, 2006), or by participating in hull adhesion in grains (Taketa et al., 2008). Finally, it is generally accepted that plant cuticle and its polymeric skeleton, cutin, are primary barriers to pathogens and that cutin monomers released by fungal cutinase are signaling molecules for both the pathogen and plants (Gilbert et al., 1996; Schweizer et al., 1996; Iwamoto et al., 2002; Yeats and Rose, 2013).The biological functions of cutin are closely controlled by its structure, which is determined by its monomer composition and by the number and position of its ester bonds. Cutin monomer composition can vary according to plant species, developmental stage (Baker et al., 1982; Peschel et al., 2007; Mintz-Oron et al., 2008), organs, and environmental stress (Espelie et al., 1979; Li-Beisson et al., 2009; Panikashvili et al., 2009; Bessire et al., 2011). Actually, cutin monomer composition determines the total number of hydroxyl (OH) groups that are potentially available for the formation of ester bonds and, therefore, the cross-linking of the polyester (Bonaventure et al., 2004; Franke et al., 2005; Peschel et al., 2007). The nonesterified OH groups enhance the hydrophilic character of the cutin polymer, increasing its elasticity (Bargel and Neinhuis, 2005).Whereas cutin monomer composition has been described extensively for different plant species, organs, and development stages, the macromolecular structure of the cutin polyester has been much less thoroughly investigated. In particular, the connectivity between the monomers is a key point for understanding the three-dimensional expansion of the polyester in relation to the polymerization process. Different approaches have been proposed to delineate the polymeric architecture of cutin. Linear dimers were identified after partial alkaline hydrolysis of tomato (Solanum lycopersicum) cutin (Osman et al., 1995). NMR and mass spectrometry analyses of oligomers released after partial depolymerization revealed primary and secondary ester linkages between cutin monomers (Graça et al., 2002; Stark and Tian, 2006) as well as covalent linkages between some cutin OH fatty acids and oligosaccharides (Tian et al., 2008). However, it has been shown that partial hydrolysis does not necessarily release all of the representative building blocks of the entire polymer (Deshmukh et al., 2003). Spectrometric analyses have also been developed for the polymer. Attenuated total reflectance (ATR)-Fourier transform infrared (FTIR) spectroscopy analyses of the methylene and carbonyl stretching vibrations allowed the estimation of an ester cross-linking index for cutin but could not differentiate the primary from the secondary ester linkages (Girard et al., 2012; Heredia-Guerrero et al., 2014). NMR studies have provided evidence of both ω- and mid-chain esters in tomato (Deshmukh et al., 2003).In this regard, tomato fruit has proved to be an interesting model for structural studies of the cutin polymer. Indeed, its astomatous cuticle can be easily isolated and is devoid of cutan. Moreover, tomato cutin composition is dominated by a monomer with two OH groups, 9(10),16-dihydroxyhexadecanoic acid (Baker et al., 1982; Osman et al., 1999; Deshmukh et al., 2003). Accordingly, the cutin monomers can be linked by either a linear (on the primary OH) or a branched (with a secondary OH) pattern. Previous studies have demonstrated that both linear and branched cross-links occur in tomato cutin. However, the relative proportion of the linear versus branched esters remains a matter of debate. Oxidation experiments have indicated that almost all of the primary cutin OH groups (94%) were involved in ester bonds, whereas only 44% of the secondary OH groups were esterified in the cutin polymer (Deas and Holloway, 1977; Kolattukudy, 1977). Conversely, partial depolymerization coupled with NMR studies of the released oligomers indicated that the branched secondary esters were the major form of tomato cutin (Graça and Lamosa, 2010). In addition, none of these studies could decipher the ester links of the glycerol OH groups.Additionally, the role of a GDSL lipase, involved in cutin polymerization, was recently reported using two different experimental approaches (Girard et al., 2012; Yeats et al., 2012). The corresponding GDSL lipase, named SlGDSL1 (Girard et al., 2012) or SlGDSL2 (Yeats et al., 2012), is now named CUTIN SYNTHASE1 (SlCUS1; Yeats et al., 2014). Different mutants affected in the expression of SlCUS1 have been generated and constitute attractive tools to delineate the structure of the cutin polymer (Girard et al., 2012; Petit et al., 2014).It has been further demonstrated that 2-monoacylglycerol (2-MAG), a putative precursor of the cutin polymer in Arabidopsis (Yang et al., 2010), can be used by a heterologously expressed SlCUS1 (Yeats et al., 2012) to produce in vitro linear oligomers in aqueous solution (Yeats et al., 2014). Nevertheless, the question of the mechanism of cutin polymerization in planta is still open. Indeed, SlCUS1 is specifically localized within the cutin matrix (i.e. a hydrophobic environment; Girard et al., 2012; Yeats et al., 2012), which could impact the acyltransferase activity of the enzyme as observed previously for lipases (Sharma et al., 2001).By coupling O-alkylation of the nonesterified OH groups of glycerol and fatty acids in an isolated cutin matrix and by further analyses of O-alkylated and nonalkylated monomers released after depolymerization, we elucidated the ester cross-link pattern of tomato cutin. We also showed at two stages of fruit development and in two different genetic backgrounds that the modulation of SlCUS1 protein level, either through RNA interference (cherry tomato ‘West Virginia 106’ [WVa106]) or mutagenesis (miniature tomato ‘Micro-Tom’), resulted in a strong alteration of the cutin ester cross-link pattern. These results give new insights into the polyester structure. In addition, while CUS1 esterification involves mostly primary OH groups in vitro (Yeats et al., 2014), our data here indicate that, in planta, deficiencies in CUS1 also affect the secondary OH group of 9(10),16-dihydroxyhexadecanoic acid and both the primary and secondary OH groups of glycerol.  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