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.
Nine new sesquiterpenes related biosynthetically to ipomeamarone, the well-known sweet potato phytoalexin, have been isolated from Ceratocystis fimbriata-infected sweet potato root tissue and the structures determined. Their biosynthetic relationship with previously identified furanoterpenes is discussed.  相似文献   

2.
Ipomeamarone accumulation in sweet potato (Ipomoea batatas) roots infected with Ceratocystis fimbriata (black rot) was decreased by one-third when roots were stored under 100 ppm ethylene. This effect of ethylene was not observed when infected tissue was also treated with benzylisothiocyanate. Ethylene treatment and long term infection were associated with the accumulation of 4-ipomeanol and 1-ipomeanol.  相似文献   

3.
All three oxygen atoms in ipomeamarone, a major stress metabolise of the sweet potato, were found to originate from molecular oxygen and not from water or the sesquiterpene precursor, farnesol.  相似文献   

4.
Characterization of major proteins in sweet potato tuberous roots   总被引:1,自引:0,他引:1  
The tuberous roots, but not other organs, of sweet potato contained large quantities of two proteins which accounted for more than 80% of the total proteins. The two proteins, tentatively named sporamins A and B, were monomeric forms with similar M,s (25 000). They were separated from each other by electrophoresis on polyacrylamide gels in a non-denaturing buffer or a buffer containing sodium dodecyl sulphate without being reduced by dithiothreitol. They were very similar to each other with respect to amino acid composition, peptide map and immunological properties. These proteins decreased in preference to other proteins during sprouting. The amino acid sequencing of the amino terminal part of sporamin A suggested that it consists of at least two molecular species with different combinations of a few amino acids.  相似文献   

5.
HgCl2 was used at up to 10 mg l–1 as an elicitor of phytoalexins in sweet potato (Ipomoea batatas (L.) Lam. cv Centennial) cell suspension cultures. Maximum stimulation of a coumarin compound was after one day of exposure using 1 mg HgCl2 l–1. The compound was identified by HPLC and GC-MS analyses as 7-hydroxycoumarin (umbelliferone).  相似文献   

6.
Two metabolites of the phytoalexin phaseollin (1) produced by cultures of the fungus Colletotrichum lindemuthianum have been assigned structures (3) and (4).  相似文献   

7.
Four sesquiterpene stress metabolites, 6-oxodendrolasin, 6-hydroxydendrolasin, 9-oxofarnesol, and 9-hydroxyfarnesol have been isolated from mercuric chloride-treated sweet potatoes. The metabolites have been synthesized and feeding studies have been carried out to determine the extent of incorporation of 14C-labelled 6-oxodendrolasin and 9-hydroxyfarnesol into ipomeamarone.  相似文献   

8.
Time course analysis of RNA contents of tissue discs after cuttingdisclosed a remarkable increase in total RNA during the first12 hr after cutting and this elevated level remained unchangedfor 48 hr. The elevated RNA level at 24 hr of incubation wasnot changed by subsequent HgCl2 treatment. The incorporationrate of the label from 3H-uridine into RNA rapidly increasedimmediately after cutting and reached a maximum at about 9 hrof incubation, then decreased sharply until 24 hr and continuedto decrease gradually thereafter. The incorporation rate at24 hr of incubation was not changed by subsequent HgCl2 treatment.The results of polyacrylamide gel electrophoresis indicatedthat bulk RNA was synthesized most actively at 9 hr of incubationthen the rate of RNA synthesis decreased gradually. (Received August 26, 1977; )  相似文献   

9.
10.
The native gibberellin A5 (GA5), as [1-3H]GA5 (3.2 Ci/mmol) was fed to seed capsules (0.58 μCi/capsule) of Pharbitis nil cv Violet at the 2-week stage of development, and its metabolism in the seeds was investigated after 43 hr. Extractable radioactivity in free GA metabolites was 38%, with 56% in GA glucosyl conjugate-like substances. Only 2.5% of the extractable radioactivity remained as [3H]GA5. Tentative identifications, based on comparisons with authentic standards after sequential chromatography on silica gel partition column → gradient-eluted C18 HPLC → isocratic-eluted C18 HPLC-radiocounting (RC), showed that [3H]GA5 was converted to at least six free GAs, GA1, GA3, GA6, GA8, GA22, GA29, a GA5 methyl ester-like metabolite, and at least twelve GA glucosyl conjugate-like substances, GA5-glucoside (GA5-G), GA5-glucosyl ester (GA5-GE), GA1-O(3)-G, GA1-O(13)-G, GA1-GE, GA3-O(3)-G, GA3-O(13)-G, GA3-GE, GA6-G or GE, GA8-O(2)-G, GA22-G or GE and GA29-O(2)-G. After lower specific activity feeds of [1,2-3H]GA5 (74 mCi/mmol; 0.1 μCi/capsule) at approximately the same stage of development, the presence of GA1, GA3, GA5, GA6, GA8 and GA29 was further confirmed by sequential (after C18 HPLC-RC) capillary gas chromatography-selected ion monitoring (GC-SIM), using six characteristic ions. However, for GA22 only a trace of the parent ion was present at the appropriate retention time.  相似文献   

11.
  • Nitrogen (N) could affect storage root growth and development of sweet potato. To manage external N concentration fluctuations, plants have developed a wide range of strategies, such as growth changes and gene expression.
  • Five sweet potato cultivars were used to analyse the functions of N in regulating storage root growth. Growth responses and physiological indicators were measured to determine the physiological changes regulated by different N concentrations. Expression profiles of related genes were analysed via microarray hybridization data and qRT‐PCR analysis to reveal the molecular mechanisms of storage root growth regulated by different N concentrations.
  • The growth responses and physiological indicators of the five cultivars were changed by N concentration. The root fresh weight of two of the sweet potato cultivars, SS19 and GS87, was higher under low N concentrations compared with the other cultivars. SS19 and GS87 were found to be having greater tolerance to low N concentration. The expression of N metabolism and storage root growth related genes was regulated by N concentration in sweet potato.
  • These results reveal that N significantly regulated storage root growth. SS19 and GS87 were more tolerant to low N concentration and produced greater storage root yield (at 30 days). Furthermore, several N response genes were involved in both N metabolism and storage root growth.
  相似文献   

12.
Sweet potato virus disease (SPVD) was common (25–30% average incidences), and farmers recognised it as an important disease, in sweet potato crops in southern Mpigi, Masaka and Rakai Districts in Uganda, but SPVD was rare in Soroti and Tororo Districts. Whiteflies, which are the vector of sweet potato chlorotic stunt crinivirus (SPCSV) a component cause of SPVD, were correspondingly common on sweet potato crops in Mpigi and rare on crops in Tororo. However, aphids, which are the vectors of sweet potato feathery mottle potyvirus (SPFMV), the other component cause of SPVD, were not found colonising sweet potato crops, and itinerant alate aphids may be the means of transmission. Different sweet potato cultivars were predominant in the different districts surveyed and four local cultivars obtained from Kanoni in S. Mpigi, where whiteflies and SPVD were common, were more resistant to SPVD than four cultivars from Busia in Tororo District, where whiteflies and SPVD were rare. However, nationally released cultivars were even more resistant than the local cultivars from Kanoni. Yield results and interviews with farmers indicated that farmers in S. Mpigi were making compromises in their choice of cultivars to grow, some key factors being SPVD susceptibility, and the yield, taste, and marketability, duration of harvest and in-ground storability of the storage roots. These compromises need to be included in an assessment of yield losses attributable to SPVD.  相似文献   

13.
Non-inhibitory concentrations of the pterocarpan phytoalexin medicarpin were completely metabolized by isolates of Fusarium solani f. sp. pisi, f. sp. cucurbitae, f. sp. phaseoli and two other F. solani isolates genetically related to f. sp. pisi during 24 hr of growth in liquid medium. The major metabolic products accumulated without significant further degradation. Medicarpin was modified at one of three adjacent carbon atoms to form either an isoflavanone derivative, a 1a-hydroxydienone derivative or 6a-hydroxymedicarpin. Whereas each isolate degraded medicarpin to one or more metabolises, the isolates varied as to which metabolise they produced. Maackiain, another pterocarpan phytoalexin, was also metabolized by all the isolates to products analogous to those formed from medicarpin. The ability to metabolize medicarpin and maackiain was not always associated with the ability to metabolize pisatin and phaseollin, two other pterocarpan phytoalexins that were degraded by several of the isolates. Tolerance of medicarpin and maackiain was similarly not always associated with tolerance to pisatin.  相似文献   

14.
The pattern of isoforms of starch branching enzyme II or Q-enzyme II in the tuberous root of sweet potato was distinct from those of other organs; altogether 7 isoforms of QEII were contained in the sweet potato plant. The QEIIf isoform, one of the two major QEII isoforms in the tuberous root, was purified to homogeneity by using a variety of HPLC columns. The purified QEIIf was a monomeric protein with a molecular mass of about 85 kDa. Western blot analysis showed that the polyclonal antibodies raised against the purified QEIIf was significantly reactive to the rice endosperm QEI, but not to the rice endosperm QEIIa. Furthermore, the sweet potato QEIIf reacted with the antiserum raised against the rice endosperm QEI, but not with that against the rice endosperm QEIIa. The results suggest that the sweet potato QEIIf is more similar to the rice endosperm QEI than to the rice endosperm QEIIa.  相似文献   

15.
Explants from stem, leaf, and storage root tissue of sweet potato (Ipomoea batatas L.) cv. Jewel, were placed on media conaining 0.1, 1.0, and 10 mg/1NAA with 0.1, 1.0, or 10 mg/1BA in a factorial experiment. Some callus formed in every treatment, but the best callus growth was on media containing 1.0 mg/1NAA and 10 mg/1BA. Roots formed over a range of treatments but were most prolific on the medium containing 1.0 mg/1NAA and 0.1 mg/1BA. Some de novo formed roots subsequently produced shoot buds in culture. Shoot formation increased the longer the original explants remained in culture without subculture. Roots could be subcultured indefinitely on agar solidified medium, but shoot regeneration did not occur after two subcultures. Shoot formation was greatest when the roots were subcultured on medium containing 1.0 mg/NAA and 0.1 mg/1BA. The cultivar Caromex followed the same regeneration pathway, but the number of shoots formed was considerably less. Regeneration in both Jewel and Caromex explants was enhanced by light.Paper No. 8292 of the Journal Series of the North Carolina Agricultural Research Service, Raleigh, NC. The use of trade names in this publication does not imply endorsement by the North Carolina Agricultural Research Service of products named, nor criticism of similar ones not mentioned.This work was done as a partial requirement for the M.S. degree at North Carolina State University.  相似文献   

16.
甘薯块根膨大过程中ATP酶活性、ATP和ABA含量的变化   总被引:18,自引:0,他引:18  
研究选用鲁薯7号和徐薯18号为材料,对甘薯块根膨大速率变化动态及其块根中可溶性碳水化合物含量、ATP含量、ATP酶活性和脱落酸(ABA)含量的变化进行了研究分析。结果表明:(1)块根膨大速率变化动态呈一双峰曲线,第一个高峰出现在栽秧后50-70d,第二个高峰出现在栽秧后120-165d;(2)块根膨大高峰期,块根中可溶性碳水化合物含量较高,ATP含量则较低;(3)块根中ATP酶活性和ABA含量变化动态与块根膨大速率变化动态相似。讨论了ATP酶和ABA在块根膨大过程中的可能作用。  相似文献   

17.
A full-length cDNA clone encoding a putative copper/zinc-superoxide dismutase (SOD) of sweet potato, Ipomoea batatas (L.) Lam. cv Tainong 57, was isolated from a cDNA library constructed in gt10 from tuber root mRNA. Nucleotide sequence analysis of this cDNA clone revealed that it comprises a complete open reading frame coding for 152 amino acid residues. The deduced amino acid sequence showed higher homology (78–86%) with the sequence of the cytosolic SOD than that of the chloroplast SOD from other plant species. The residues required for coordinating copper and zinc are conserved as they are among all reported Cu/Zn-SOD sequences. In addition, it lacks recognizable plastic or mitochondrial targeting sequences. These data suggest that the isolated sweet potato clone encodes a cytosolic Cu/Zn-SOD.  相似文献   

18.
Three hundred and ninety‐four sweet potato accessions from Latin America and East Africa were screened by polymerase chain reaction (PCR) for the presence of begomoviruses, and 46 were found to be positive. All were symptomless in sweet potato and generated leaf curling and/or chlorosis in Ipomoea setosa. The five most divergent isolates, based on complete genome sequences, were used to study interactions with Sweet potato chlorotic stunt virus (SPCSV), known to cause synergistic diseases with other viruses. Co‐infections led to increased titres of begomoviruses and decreased titres of SPCSV in all cases, although the extent of the changes varied notably between begomovirus isolates. Symptoms of leaf curling only developed temporarily in combination with isolate StV1 and coincided with the presence of the highest begomovirus concentrations in the plant. Small interfering RNA (siRNA) sequence analysis revealed that co‐infection of SPCSV with isolate StV1 led to relatively increased siRNA targeting of the central part of the SPCSV genome and a reduction in targeting of the genomic ends, but no changes to the targeting of StV1 relative to single infection of either virus. These changes were not observed in the interaction between SPCSV and the RNA virus Sweet potato feathery mottle virus (genus Potyvirus), implying specific effects of begomoviruses on RNA silencing of SPCSV in dually infected plants. Infection in RNase3‐expressing transgenic plants showed that this protein was sufficient to mediate this synergistic interaction with DNA viruses, similar to RNA viruses, but exposed distinct effects on RNA silencing when RNase3 was expressed from its native virus, or constitutively from a transgene, despite a similar pathogenic outcome.  相似文献   

19.
A large increase in the activity of hydroxycinnamyl CoA:quinate hydroxycinnamyl transferase (CQT) occurred in potatoes stored at 0 and 2° and such an increase was prevented by storage at either 5 or 10°. The increase was most rapid in potatoes stored at 0° where it reached a maximum after 28 days and then declined slowly during storage for up to 6 months. Accompanying these changes in CQT were transitory increases in p-coumarate CoA ligase and PAL which occured during the first few weeks of storage at 0° and during this period there was nearly a two fold increase in the chlorogenic acid content of the tissue. The increase in chlorogenic acid did not occur at 10° when the increases in PAL, ligase and CQT were also prevented. The increase in CQT was reversed when tubers stored at 0° for 14 days were returned to 10° and this warming up period prevented further increase in CQT on return to 0°. The increase in CQT at 0° was prevented if the air in the storageatmosphere was replaced by N2, 1 % O2 or 10–15% CO2. Similar increases in CQT, ligase and chlorogenic acid occurred in sweet potatoes stored at 7.5° but were prevented by storage at 15°. The role of PAL, ligase and CQT in the control of chlorogenic acid accumulation in these commodities and the significance of changes in their activities in relation to physiological changes at low temperatures are discussed.  相似文献   

20.
The coding region of copper/zinc-superoxide dismutase (Cu/Zn-SOD) cDNA from sweet potato, Ipomoea batatas (L.) Lam. cv. Tainong 57, was introduced into an expression vector, pET-20b(+). The Cu/Zn-SOD purified by His-tagged technique showed two active forms (dimer and monomer). The amount of proteins of dimer and monomer appeared to be equal, but the activity of dimeric form was seven times higher than that of monomeric form. The enzyme was dissociated into monomer by imidazole buffer above 1.0 M, acidic pH (below 3.0), or SDS (above 1%). The enzyme is quite stable. The enzyme activity is not affected at 85 °C for 20 min, in alkali pH 11.2, or in 0.1 M EDTA and also quite resistant to proteolytic attack. Dimer is more stable than monomer. The thermal inactivation rate constant k dcalculated for the monomer at 85 °C was 0.029 min-1 and the half-life for inactivation was about 28 min. In contrast, there is no significant change of dimer activity after 40 min at 85 °C. The enzyme dimer and monomer retained 83% and 58% of original activity, respectively, after 3 h incubation with trypsin at 37 °C, while those retained 100% and 31% of original activity with chymotrypsin under the same condition. These results suggest subunit interaction might change the enzyme conformation and greatly improve the catalytic activity and stability of the enzyme. It is also possible that the intersubunit contacts stabilize a particular optimal conformation of the protein or the dimeric structure enhances catalytic activity by increasing the electrostatic steering of substrate into the active site.  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