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 0 毫秒
1.
In a study of relationships among selected cyst-forming and noncyst-forming species of Heteroderoidea, combined sequences comprised of DNA from part of the conserved 18S ribosomal RNA gene (rDNA) plus the complete ITS rDNA segment were more similar to analyses based on the ITS data alone than to analyses based on the 18S data alone. One of the two noncyst-forming species, Ekphymatodera thomasoni, grouped with cyst-forming species of Heteroderoidea. Bilobodera flexa, also a noncyst-forming species, was separated from all the other taxa by a long branch. Afenestrata koreana, with a weakly sclerotized cyst, grouped closely with H. bifenestra. These observations suggest that phylogenetic analyses using molecular data may aid in our understanding of the evolution of cyst formation in nematodes, including the possibility of secondary loss. The usefulness of molecular phylogenetic analyses in nematodes may depend more on the particular selection of taxa than on mere addition of data from additional genes.  相似文献   

2.
Evolutionary relationships based on ribosomal DNA (rDNA) sequence data for a previously unknown species of Globodera from Portugal, Punctodera chalcoensis from Mexico, and P. punctata from Estonia, plus previously published sequences, support the following relationships: (((Cactodera weissi, G. artemisiae, C. milleri), ((G. sp. Bouro, G. sp. Canha, G. sp. Ladoeiro), ((G. pallida, G. rostochiensis), (P. chalcoensis, P. punctata)))), Heterodera avenae). Globodera sp. from Portugal, which can be confused with potato cyst nematodes by phytosanitary services when the identification is based only on morphological characters, is clearly different based on our molecular data. In addition, the rDNA data show the Globodera sp. to be only distantly related to other European Globodera species that parasitize Asteraceae. Punctodera chalcoensis and P. punctata form a sister clade to the G. pallida + G. rostochiensis clade.  相似文献   

3.
Ribosomal DNA (rDNA) sequence data were compared for five species of Globodera, including G. rostochiensis, G. pallida, G. virginiae, and two undescribed Globodera isolates from Mexico collected from weed species and maintained on Solanum dulcamara. The rDNA comparisons included both internal transcribed spacers (ITS1 and ITS2), the 5.8S rRNA gene, and small portions of the 3'' end of the 18S gene and the 5'' end of the 28S gene. Phylogenetic analysis of the rDNA sequence data indicated that the two potato cyst nematodes, G. pallida and especially G. rostochiensis, are closely related to the Mexican isolates, whereas G. virginiae is relatively dissimilar to the others and more distantly related. The data are consistent with the thesis that Mexico is the center of origin for the potato cyst nematodes.  相似文献   

4.
Phylogenetic analysis of new ribosomal DNA (rDNA) data for Heterodera mediterranea, H. hordecalis, H. carotae, and H. fici from Italy and H. ciceri from Syria, along with published data for other species, showed high bootstrap support for the following relationships: (((((H. carotae H. cruciferae) H. goettingiana) (((H. trifolii H. ciceri) H. mediterranea) ((H. avenae H. latipons) H. fici))) (Cactodera betulae H. hordecalis)) (Globodera rostochiensis G. pallida)). The rDNA sequence data were for the two internal transcribed spacers (ITS1 and ITS2) plus the 5.8S gene between them. These inferred relationships support the classic ''''Goettingiana Group'''' of H. carotae, H. cruciferae, and H. goettingiana. A clade comprised of Cactodera betulae and H. hordecalis is only distantly related to the other species in the analysis.  相似文献   

5.
Six geographic isolates of Heterodera avenae, including two isolates each from Sweden, Australia, and the United States, were compared on the basis of 2-D PAGE protein patterns and the complete DNA sequence for the two internal transcribed ribosomal DNA spacers (rDNA ITS1 and ITS2) and the 5.8S rRNA gene. The protein pattern data and rDNA ITS sequence data both indicated that the Swedish Gotland strain of H. avenae differed markedly from the rest of the isolates. Protein patterns for the Australia isolates differed more from a Swedish strict H. avenae isolate and isolates from Oregon and Idaho, than the two U.S. isolates and the Swedish strict H. avenae isolate differed from each other. Except for the Gotland strain isolate, the rDNA ITS sequences were highly conserved among all of the H. avenae isolates, just as we earlier found them to be conserved among species of the schachtii group of Heterodera.  相似文献   

6.
Cactodera salina n. sp. (Heteroderinae) is described from roots of the estuary plant Salicornia bigelovii (Chenopodiaceae), in Puerto Pefiasco, Sonora, Mexico, at the northern tip of the Sea of Cortez. The halophyte host is grown experimentally for oilseed in plots flooded daily with seawater. Infected plants appear to be adversely affected by C. salina relative to plants in noninfested plots. Cactodera salina extends the morphological limits of the genus. Females and cysts have a very small or absent terminal cone and deep cuticular folds in a zigzag pattern more typical of Heterodera and Globodera than of Cactodera spp. Many Cactodera spp. have a tuberculate egg surface, whereas C. salina shares the character of a smooth egg with C. amaranthi, C. weissi, and C. acnidae. Only C. milleri and C. acnidae have larger cysts than C. salina. Face patterns of males and second-stage juveniles, as viewed with scanning electron microscopy, reveal the full complement of six lip sectors as in other Cactodera spp. Circumfenestrae of C. salina are typical for the genus.  相似文献   

7.
Genetic analyses using DNA sequences of nuclear ribosomal DNA ITS1 were conducted to determine the extent of genetic variation within and among Longidorus and Xiphinema species. DNA sequences were obtained from samples collected from Arkansas, California and Australia as well as 4 Xiphinema DNA sequences from GenBank. The sequences of the ITS1 region including the 3'' end of the 18S rDNA gene and the 5'' end of the 5.8S rDNA gene ranged from 1020 bp to 1244 bp for the 9 Longidorus species, and from 870 bp to 1354 bp for the 7 Xiphinema species. Nucleotide frequencies were: A = 25.5%, C = 21.0%, G = 26.4%, and T = 27.1%. Genetic variation between the two genera had a maximum divergence of 38.6% between X. chambersi and L. crassus. Genetic variation among Xiphinema species ranged from 3.8% between X. diversicaudatum and X. bakeri to 29.9% between X. chambersi and X. italiae. Within Longidorus, genetic variation ranged from 8.9% between L. crassus and L. grandis to 32.4% between L. fragilis and L. diadecturus. Intraspecific genetic variation in X. americanum sensu lato ranged from 0.3% to 1.9%, while genetic variation in L. diadecturus had 0.8% and L. biformis ranged from 0.6% to 10.9%. Identical sequences were obtained between the two populations of L. grandis, and between the two populations of X. bakeri. Phylogenetic analyses based on the ITS1 DNA sequence data were conducted on each genus separately using both maximum parsimony and maximum likelihood analysis. Among the Longidorus taxa, 4 subgroups are supported: L. grandis, L. crassus, and L. elongatus are in one cluster; L. biformis and L. paralongicaudatus are in a second cluster; L. fragilis and L. breviannulatus are in a third cluster; and L. diadecturus is in a fourth cluster. Among the Xiphinema taxa, 3 subgroups are supported: X. americanum with X. chambersi, X. bakeri with X. diversicaudatum, and X. italiae and X. vuittenezi forming a sister group with X. index. The relationships observed in this study correspond to previous genera and species defined by morphology.  相似文献   

8.
9.
Maximum likelihood trees produced from 18S rDNA sequences separated 14 Xiphinema and five Xiphidorus nematode species from Brazil into distinct groups that concurred with their current morphological taxonomic status. Species belonging to the X. americanum group (X. brevicolle, X. diffusum, X. oxycaudatum, and X. peruvianum) formed a single group that was clearly separated from the other Xiphinema species. As with previous taxonomic studies that noted only minor morphological differences between putative X. americanum group species, separation of these species based upon 18S rDNA sequences was inconclusive. Thus it is probable that instead of comprising distinct species, the X. americanum group may in fact represent numerous morphotypes with large inter- and intra- population morphological variability that may be environmentally driven. Within the cluster representing non X. americanum group species, there was little statistical support to clearly separate species. However, three subgroups, comprising (i) the X. setariae/vulgare complex, (ii) X. ifacolum and X. paritaliae, and (iii) X. brasiliense and X. ensiculiferum were well resolved.  相似文献   

10.
A morphometric evaluation of second-stage juveniles (J2), males, females, cysts, and eggs of several isolates of the tobacco cyst nematode (TCN) complex, Globodera tabacum tabacum (GTT), G. t. virginiae (GTV), and G. t. solanacearum (GTS) is presented. Morphometrics of eggs, J2, and males are considerably less variable than of females and cysts. No measurements of eggs and J2 are useful for identification of the three subspecies. Distance from the median bulb and excretory pore to the head end in J2 and males is quite stable. Stylet knob width of males is useful for identifying GTV isolates and tail length in separating males of GTT isolates from GTV and GTS. Body length/width (L/W) ratio of females and cysts discriminates GTT from GTV and GTS; stylet knob width is an auxiliary character for identifying GTV. This subspecies complex has a continuum of values for the other characters. Data suggest a close relationship between GTV and GTS, which also occur in close proximity in Virginia.  相似文献   

11.
Fructose-bisphosphate aldolase (EC 4.1.2.13) is a key enzyme in glycolysis. We have characterized full-length coding sequences for aldolase genes from the cyst nematodes Heterodera glycines and Globodera rostochiensis, the first for any plant-parasitic nematode. Nucleotide homology is high (83% identity), and the respective sequences encode 40 kDa proteins with 89% amino acid identity. Genomic sequences contain six introns located at identical positions in both genes. Intron 4 in the H. glycines gene is >500 bp. Partial genomic sequences determined for seven other cyst nematode species reveal that the large fourth intron is characteristic of Heterodera but not Globodera aldolase genes. Total aldolase-like specific activity in homogenates from H. glycines was 2-fold lower than in either Caenorhabditis elegans or Panagrellus redivivus (P = 0.001). Activity in H. glycines samples was higher in juvenile stages than in adults (P = 0.003). Heterodera glycines aldolase has Km = 41 µM and is inhibited by treatment with carboxypeptidase A or sodium borohydride.  相似文献   

12.
Globodera rostochiensis and G. pallida responded similarly to hatch stimulation by potato root leachate, but proportionally more second-stage juveniles (J2s) of G. rostochiensis hatched than of G. pallida in response to picrolonic acid, sodium thiocyanate, alpha-solanine, and alpha-chaconine. Fractionation of the potato root leachate identified hatching factors with species-selective (active toward both species but stimulating greater hatch of one species than the other), -specific (active toward only one species), and -neutral (equally active toward both species) activities. In a comparison of two populations of each of the two potato cyst nematode (PCN) species, however, greater similarity in response to the individual hatching factors was observed among populations of different species produced under the same conditions than among different populations of the same PCN species. Smaller numbers of species-specific and species-selective hatching factor stimulants and hatching inhibitors than of hatching factors were resolved. In a study to determine whether the different hatching responses of the two species to the same root leachate were associated with different ratios of species-selective and species-specific hatching factors, G. rostochiensis pathotype Ro1 exhibited greater hatch than did G. pallida pathotype Pa2/3 in response to leachate from older plants (more than 38 days old), while G. pallida exhibited greater hatch in response to leachate from younger plants (less than 38 days old); the response of G. pallida pathotype Pal with respect to plant age was intermediate between the other two populations. Combined molecular exclusion-ion exchange chromatography of the root leachates from plants of different ages revealed an increase in the proportion of G. rostochiensis-specific and -selective hatching factors as the plants aged.  相似文献   

13.
Chamaecrista belongs to subtribe Cassiinae (Caesalpinioideae), and it comprises over 330 species, divided into six sections. The section Xerocalyx has been subjected to a profound taxonomic shuffling over the years. Therefore, we conducted a phylogenetic analysis using a cpDNA trnE-trnT intergenic spacer and nrDNA ITS/5.8S sequences from Cassiinae taxa, in an attempt to elucidate the relationships within this section from Chamaecrista. The tree topology was congruent between the two data sets studied in which the monophyly of the genus Chamaecrista was strongly supported. Our analyses reinforce that new sectional boundaries must be defined in the Chamaecrista genus, especially the inclusion of sections Caliciopsis and Xerocalyx in sect. Chamaecrista, considered here paraphyletic. The section Xerocalyx was strongly supported as monophyletic; however, the current data did not show C. ramosa (microphyllous) and C. desvauxii (macrophyllous) and their respective varieties in distinct clades, suggesting that speciation events are still ongoing in these specimens.  相似文献   

14.
Internal transcribed spacer 1 sequences were used to infer phylogenetic relationships among 8 of the 9 described species and one putative species of the entomopathogenic nematode genus Heterorhabditis. Sequences were aligned and optimized based on pairwise genetic distance and parsimony criteria and subjected to a variety of sequence alignment parameters. Phylogenetic trees were constructed with maximum parsimony, cladistic, distance, and maximum likelihood algorithms. Our results gave strong support for four pairs of sister species, while relationships between these pairs also were resolved but less well supported. The ITS1 region of the nuclear ribosomal repeat was a reliable source of homologous characters for resolving relationships between closely related taxa but provided more tenuous resolution among more divergent lineages. A high degree of sequence identity and lack of autapomorphic characters suggest that sister species pairs within three distinct lineages may be mutually conspecific. Application of these molecular data and current morphological knowledge to the delimitation of species is hindered by an incomplete understanding of their variability in natural populations.  相似文献   

15.
Three genes in the major sperm protein (MSP) gene family from the potato cyst nematode Globodera rostochiensis were cloned and sequenced. In contrast to the absence of introns in Caenorhabditis elegans MSP genes, these genes in G. rostochiensis contained a 57 nucleotide intron, with normal exon-intron boundaries, in the same relative location as the intron in Onchocerca volvulus. The MSP genes of G. rostochiensis had putative CAAT, TATA, and polyadenylation signals. The predicted G. rostochiensis MSP gene product is 126 amino acids long, one residue shorter than the products in the other species. The comparison of MSP amino acid sequences from four diverse nematode species suggests that O. volvulus, Ascaris suum, and C. elegans may be more closely related to each other than they are to G. rostochiensis.  相似文献   

16.
Phylogenies were inferred from nearly complete small subunit (SSU) 18S rDNA sequences of 12 species of Meloidogyne and 4 outgroup taxa (Globodera pallida, Nacobbus abberans, Subanguina radicicola, and Zygotylenchus guevarai). Alignments were generated manually from a secondary structure model, and computationally using ClustalX and Treealign. Trees were constructed using distance, parsimony, and likelihood algorithms in PAUP* 4.0b4a. Obtained tree topologies were stable across algorithms and alignments, supporting 3 clades: clade I = [M. incognita (M. javanica, M. arenaria)]; clade II = M. duytsi and M. maritima in an unresolved trichotomy with (M. hapla, M. microtyla); and clade III = (M. exigua (M. graminicola, M. chitwoodi)). Monophyly of [(clade I, clade II) clade III] was given maximal bootstrap support (mbs). M. artiellia was always a sister taxon to this joint clade, while M. ichinohei was consistently placed with mbs as a basal taxon within the genus. Affinities with the outgroup taxa remain unclear, although G. pallida and S. radicicola were never placed as closest relatives of Meloidogyne. Our results show that SSU sequence data are useful in addressing deeper phylogeny within Meloidogyne, and that both M. ichinohei and M. artiellia are credible outgroups for phylogenetic analysis of speciations among the major species.  相似文献   

17.
The establishment of Globodera rostochiensis Rol populations was examined under greenhouse conditions. The probability of G. rostochiensis population establishment was calculated from the number of plants that produced new cysts with viable eggs following inoculation with various numbers of eggs of different ages. Probability of population establishment was positively correlated with inoculum density but was not affected by the age of eggs used in these experiments. The probability of G. rostochiensis establishment ranged from 5% at densities of 2 eggs/pot to 100% at densities of 25 eggs/pot or greater. At densities of 3 eggs/pot and beyond, there was no correlation between inoculum density and the number of viable eggs/new cyst. Also, the number of plants that produced new cysts was a function of inoculum density and not age of eggs. Juveniles from eggs 1 year old or older were equally as infective as were those from eggs in newly developed cysts (4 months old).  相似文献   

18.
Detailed morphological comparisons with light and scanning electron microscopy were made of white females and cysts of several isolates of Globodera tabacum sspp. tabacum (GTT), virginiae (GTV), and solanacearum (GTS). Observations focused on body shape, anterior region including head shape, lip pattern, stylet morphology, and the terminal area in females; and body shape and terminal area of cysts. The most useful characters to separate the three subspecies were forms of the female body, cyst, stylet knobs, tail region, perineal tubercles, anal-fenestral ridge patterns, and the distinctiveness of the anus. GTT is characterized by having round females and cysts, sharply back sloped stylet knobs, clumped perineal tubercles in the vulval region, tight parallel ridges in the cyst anal-fenestral region, and a uniformly conoid tail region. GTV is characterized by its ovoid to ellipsoid female and cyst shape, the "Dutch shoe" shape of the dorsal stylet knob, the more dispersed perineal tubercles, a maze-like pattern of ridges in the anal-fenestral region, and an indistinct anus. GTS is characterized by its ovoid to ellipsoid female and cyst shape, moderately backward sloped stylet knobs, more widely separated ridges, a distinct anus, and a usually crescent shaped tail region. Much variability in shape and patterns is visible among all the isolates of the different subspecies. Tubercles in the neck, as well as bullae, are reported, and their taxonomic value is discussed.  相似文献   

19.
The unusual arrangement of the 5S ribosomal gene within the intergenic sequence (IGS) of the ribosomal cistron, previously reported for Meloidogyne arenaria, was also found in the ribosomal DNA of two other economically important species of tropical root-knot nematodes, M, incognita and M. javanica. This arrangement also was found in M. hapla, which is important in temperate regions, and M. mayaguensis, a virulent species of concern in West Africa. Amplification of the region between the 5S and 18S genes by PCR yielded products of three different sizes such that M. mayaguensis could be readily differentiated from the other species in this study. This product can be amplified from single juveniles, females, or egg masses. The sequences obtained in this region for one line of each of M. incognita, M. arenaria, and M. javanica were very similar, reflecting the close relationships of these lineages. The M. mayaguensis sequence for this region had a number of small deletions and insertions of various sizes, including possible sequence duplications.  相似文献   

20.
Morphological comparisons with light microscopy and scanning electron microscopy were made among second-stage juveniles (J2) and males of several isolates of the three subspecies of the tobacco cyst nematode complex, Globodera tabacum sspp. tabacum, virginiae, and solanacearum. Observations focused on the anterior region, (including head shape, lip pattern, and stylet morphology) and the tail region (including tail shape in J2 and spicules in males). The three subspecies could not be separated on the basis of any of these characters.  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