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 31 毫秒
1.
Summary Using disomic chromosome substitution lines based on the susceptible wheat cultivar Chinese Spring, loose smut resistance of wheat cultivars Hope and Thatcher was shown to be conferred in each case by a single dominant major gene carried on chromosome 7 A (Hope) or 7 B (Thatcher). Partial resistance was determined by genes on an additional eight Hope or seven Thatcher chromosomes, and similarities were evident between the partial resistance genotypes ofHope and Thatcher. Chinese Spring exhibited a mean infection value of approximately 50%, indicating a significant level of partial resistance, which was found to be due, in part, to genes on the homoeologous chromosome arms 1 As, 1 Es and 1 Ds, and to cytoplasmic genes. Substitution of the Chinese Spring nucleus into the cytoplasm of Aegilops squarrosa, Ae. variabilis or Ae. mutica resulted in increased susceptibility to Ustilago tritici. Several alloplasmic lines of the resistant wheat cultivars Selkirk and Chris exhibited race-specific susceptibility to U. tritici.  相似文献   

2.
Many genes have been located in wheat chromosomes, yet little is known about the location of genes for resistance to Ustilago tritici, which causes loose smut. Crosses were made between the loose smut susceptible alien substitution lines Cadet 6Ag(6A) and Rescue 6Ag(6A) (lines in which Agropyron chromosome 6 is substituted by wheat chromosome 6A) and four cultivars resistant to U. tritici race T19: Cadet, Kota, Thatcher and TD18. The segregating progeny were tested for reaction to race T19 and for the level of binding with a monoclonal antibody specific to a chromosome 6A-coded seed protein. The antibody, which does not bind to seed protein extracts in the absence of the 6A chromosome, was used as a chromosome marker. An association was established between resistance to race T19 and the presence of chromosome 6A for each of the cultivars tested, indicating that resistance to race T19 resides in chromosome 6A. Ustilago tritici race T19 resistance in Cadet appears to be located in the short arm of chromosome 6A, based on the evaluation of the Cadet 6A long ditelosomic stock, which was susceptible, and the Cadet 6A-short: 6-Agropyron- short alien translocation stock, which was resistant.  相似文献   

3.
Summary Thirteen maize (Zea mays L.) populations including five adapted, five adapted x exotic, two composites of adapted and exotic, and one exotic selected for adaptability were crossed in a diallel mating system. The parents and 78 crosses and nine check hybrids were evaluated for grain yield and plant height in five environments. The Gardner-Eberhart model Analysis II indicated that additive and nonadditive gene effects accounted for 60 and 40% of the total variation among populations, respectively, for grain yield and 86% and 14% of the total variation, respectively, for plant height. Components of heterosis were significant in the combined analysis for both traits. Adapted Corn Belt populations tended to have higher performance in crosses and greater values of variety heterosis than 50% adapted populations. Nebraska Elite Composite, Corn Belt x Mexican, and Corn Belt x Brazilian showed high mean yields in crosses, however, they were not among those with high estimates of variety heterosis. One exotic population (Tuxpeno x Antigua Grupo 2) and three adapted populations [307 Composite, NB(S1)C-3, and NK(S1)C-3] might be combined together to form a high-yielding population. It may be possible to synthesize two useful populations for reciprocal recurrent selection by grouping Tuxpeno x Antiqua Grupo 2, NB(S1)C-3, and NS(FS)LFW-8 into one population and NK(S1)C-3, KrugxTabloncillo, and 307 Composite in the other one.Contribution from the Department of Agronomy, University of Nebraska, Lincoln, NE 68583. Published as Paper No. 8011. Journal Series, Nebr. Agric. Exp. Station. Research was conducted under Project 12-049  相似文献   

4.
Recombinant Penicillium citrinum -1,2-mannosidase, expressed in Aspergillus oryzae, was employed to carry out regioselective synthesis of -d-mannopyranosyl-(12)-d-mannose. Yields (w/w) of 16.68% disaccharide, 3.07% trisaccharide and 0.48% tetrasaccharide were obtained, with 12 linkages present at 98.5% of the total linkages formed. Non-specific -mannosidase from almond was highly efficient in reverse hydrolysis and oligosaccharide yields of 45–50% were achieved. The products of the almond mannosidase were a mixture of disaccharides (30.75%, w/w), trisaccharides (12.26%, w/w) and tetrasaccharides (1.89%, w/w) with 12, 13 and 16 isomers. -1,2-linkage specific mannosidase from P. citrinum and -1,6-linkage-specific mannosidase from Aspergillus phoenicis were used in combination to hydrolyse the respective linkages from the mixture of isomers, resulting in -d-mannopyranosyl-(13)-d-mannose in 86.4% purity. The synthesised oligosaccharides can potentially inhibit the adhesion of pathogens by acting as "decoys" of receptors of type-1 fimbriae carried by enterobacteria.  相似文献   

5.
The presence of a newly formed primary cell wall was shown to be required for attachment and subsequent transformation of tobacco leaf protoplasts by Agrobacterium tumefaciens in cocultivation experiments. In these experiments both protoplasts at different stages after their isolation and cell-wall inhibitors were used. The specificity of Agrobacterium attachment was shown by using other kinds of bacteria that did not attach. By diminishing the concentration of divalent cations using ethylenediaminetetraacetic acid, neither attachment nor transformation was found; however, when more specifically the Ca2+concentration was lowered by ethylene glycol-bis (-aminoethyl ether)-N,N,N,N-tetraacetic acid, both phenomena occurred. Commercial lectins had no effect on binding, but this observation does not exclude the involvement of other lectins. Protoplasts isolated from various crown-gall callus tissues also developed binding sites, but when they were at the stage of dividing cells, attachment of agrobacteria was no longer observed. In this respect, cells from protoplasts of normal tobacco leaves behaved differently. Even 16 d after protoplast isolation, the dividing cells were still able to bind A. tumefaciens, while transformation was not detected. For transformation of 3-d-old tobacco protoplasts, a minimal co-cultivation period of 24 h was required, while optimal attachment took place within 5 h. It is concluded that the primary cell wall was sufficiently well formed that certain functional receptor molecules were available for attachment of Agrobacterium as the first step of a multistep process leading to the transformation of cells. The expression of bacterial functions required for attachment, moreover, was independent of the presence of Ti-plasmid.Abbreviations ConA concanavalin A - CW calcofluor white - EDTA ethylenediaminetetraacetic acid - EGTA ethylene glycol-bis (-aminoethyl ether)-N,N,N,N-tetraacetic acid - -Man -methyl-d-mannoside  相似文献   

6.
Inheritance of resistance to blackmold, a disease of ripe tomato fruit caused byAlternaria alternata, was studied in two interspecific crosses. The parents, F1 and F2 generations of a cross between the susceptibleLycopersicon esculentum Mill. cultivar Hunt 100 and the resistantL. Cheesmanii f.typicum Riley accession LA 422, and the parents, F1, F2, F3, and BC1 P2 generations of a cross between the susceptibleL. Esculentum cv. VF 145B-7879 and LA 422 were evaluated. The following disease evaluation traits were used: symptom rating (a symptom severity rating based on visual evaluation of lesions), diseased fruit (the number of diseased fruits divided by the total number of fruit scored), and lesion size (a function derived from the actual lesion diameter). Generation means analysis was used to determine gene action. The data of the Hunt 100 × LA 422 cross fit an additive-dominance model for all three traits. The VF 145B-7879 × LA 422 cross data best fit a model that included the additive × additive and additive × dominance interaction components for the trait diseased fruit, whereas higher-order epistatic models would have to be invoked to fit the data for the traits symptom rating and lesion size. A minimum of one gene segregated for all three traits. Broad-sense heritability estimates ranged from 0.09 to 0.16 for all three traits, indicating that selection for improved resistance to blackmold will require selection on a family performance basis.  相似文献   

7.
We have taken a systematic genetic approach to study the potential role of glutathione metabolism in aluminum (Al) toxicity and resistance, using disruption mutants available in Saccharomyces cerevisiae. Yeast disruption mutants defective in phospholipid hydroperoxide glutathione peroxidases (PHGPX; phgpx1 , phgpx2 , and phgpx3), were tested for their sensitivity to Al. The triple mutant, phgpx1 /2/3, was more sensitive to Al (55% reduction in growth at 300 M Al) than any single phgpx mutant, indicating that the PHGPX genes may collectively contribute to Al resistance. The hypersensitivity of phgpx3 to Al was overcome by complementation with PHGPX3, and all PHGPX genes showed increased expression in response to Al in the wild-type strain (YPH250), with maximum induction of approximately 2.5-fold for PHGPX3. Both phgpx3 and phgpx1/2/3 mutants were sensitive to oxidative stress (exposure to H2O2 or diamide). Lipid peroxidation was also increased in the phgpx1/2/3 mutant compared to the parental strain. Disruption mutants defective in genes for glutathione S-transferases (GSTs) (gtt1 and gtt2), glutathione biosynthesis (gsh1 and gsh2), glutathione reductase (glr1) and a glutathione transporter (opt1) did not show hypersensitivity to Al relative to the parental strain BY4741. Interestingly, a strain deleted for URE2, a gene which encodes a prion precursor with homology to GSTs, also showed hypersensitivity to Al. The hypersensitivity of the ure2 mutant could be overcome by complementation with URE2. Expression of URE2 in the parental strain increased approximately 2-fold in response to exposure to 100 M Al. Intracellular oxidation levels in the ure2 mutant showed a 2-fold (non-stressed) and 3-fold (when exposed-to 2 mM H2O2) increase compared to BY4741; however, the ure2 mutant showed no change in lipid peroxidation compared to the control. The phgpx1/2/3 and ure2 mutants both showed increased accumulation of Al. These findings suggest the involvement of PHGPX genes and a novel role of URE2 in Al toxicity/resistance in S. cerevisiae.Communicated by D.Y. Thomas  相似文献   

8.
Stimulation of 1-adrenoceptors (AR) during ischaemia in the rat heart by exogenous phenylephrine exacerbates reperfusion arrhythmias, an effect apparently mediated by the 1A-AR subtype. We tested whether 1A-AR stimulation byendogenous catecholamines, released during ischaemia, could modulate reperfusion arrhythmias, using as pharmacological tools the selective 1A-AR antagonists abanoquil (UK52046) and WB4101. Isolated rat hearts (n=12/group) were subjected to dual coronary perfusion. After 15 min of aerobic perfusion of both coronary beds, abanoquil or WB4101 was infused selectively into the left coronary bed (LCB) for 5 min. The LCB was then subjected to 10 min of zero-flow ischaemia and 5 min of reperfusion. Effects on PR interval, width of the ventricular complex (QRST90) and reperfusion arrhythmias were assessed. Abanoquil at concentrations of 0.03, 0.1 and 0.3 M tended to reduce the incidence of reperfusion-induced ventricular fibrillation (VF) in a dose-dependent manner from 75% in controls to 58, 33 and 25%, but this effects did not achieve statistical significance. Similarly, WB4101 at 0.1, 0.3 and 1 M also tended to reduce VF incidence from 67% in controls to 67, 42% and 33% (NS). The incidence of ventricular tachycardia (VT) was 100% in all groups and ECG parameters were not altered significantly by either drug. These results suggest that, in this denervated isolated heart preparation, 1A-AR stimulation during ischaemia by endogenous catecholamines does not significantly modulate reperfusion arrhythmias.  相似文献   

9.
Summary One hundred and twenty-two varieties, lines and wild accessions of Lycopersicon were screened under three different regimes during the autumn/winter season of 1982–83 and 1983–84 for resistance to tomato leaf curl virus (TLCV). L. hirsutum f. glabratum (B6013) and L. hirsutum f. typicum (A1904) proved to be highly resistant to TLCV in all three environments. Various accessions of L. peruvianum were also highly resistant. L. pimpinellifolium (A1921) exhibited no TLCV symptoms within 90 days. Of the cultivated varieties, Acc 99 exhibited the minimim score for susceptibility; AC 142, Collection No. 2, Kalyanpur Angurlata and HS 101 had a low rating for virus incidence. The inheritance of resistance was studied in the interspecific crosses between a TLCV resistant line of L. pimpinellifolium (A1921) and five (HS 101, HS 102, HS 110, Pusa Ruby and Punjab Chhuhara) susceptible cultivars of L. esculentum. Parents, F1, F2 and backcross progenies were artificially inoculated with local strains of TLCV using vector the viruliferious whitefly, Bemisia tabaci (Genn.). Data indicated that the resistance of L. pimpinellifolium (A 1921) is monogenic and incompletely dominant over susceptibility.  相似文献   

10.
Two marihuana constituents, cannabidiol (1) and cannabidiolic acid (4) were each converted mainly to two metabolites using tissue segments of Pinellia ternata tuber. The structures of the metabolites formed from 1 were determined to be 1-O-D-glucopyranoside and 1-O-D-diglucopyranoside by 1H nmr, 13C nmr and fabms. Those from 4 were determined as 4-O-D-glucopyranoside and 10-hydroxyl 4-O-D-glucopyranoside. In time course experiments, 1 was absorbed rapidly by the tissues and glucosylated. Hydroxylation subsequent to the glucosylation occurred at the pentyl group in 4.1 For Part 24, see Taura F, Morimoto S, Shoyama Y, Mechoulam R (1995) J Am Chem Soc 117:9766–9767  相似文献   

11.
Novel O-serotypes were revealed among Pseudomonas syringae pv. garcae strains by using a set of mouse monoclonal antibodies specific to the lipopolysaccharide O-polysaccharide. Structural studies showed that the O-polysaccharide of P. syringae pv. garcae NCPPB 2708 is a hitherto unknown linear L-rhamnan lacking strict regularity and having two oligosaccharide repeating units I and II, which differ in the position of substitution in one of the rhamnose residues and have the following structures: I:3)--L Rha (12)-- L Rha (12)--L-Rha-(13)--L Rha (1;II: 2)--L-Rha-(13) -L-Rha-(12)--L-Rha-(13)--L Rha (1.The branched O-polysaccharides of P. syringae pv. garcae ICMP 8047 and NCPPB 588T have the same L-rhamnan backbone with repeating units I and II and a lateral chain of 14)- or 13)-linked residues of 3-acetamido-3,6-dideoxy-D-galactose (D-Fuc3NAc). Several monoclonal antibody epitopes associated with the L-rhamnan backbone or the lateral -D-Fuc3NAc residues were characterized.Translated from Mikrobiologiya, Vol. 73, No. 6, 2004, pp. 777–789.Original Russian Text Copyright © 2004 by Ovod, Zdorovenko, Shashkov, Kocharova, Knirel.  相似文献   

12.
Molecular markers for the crown rust resistance genes Pc38, Pc39, and Pc48 in cultivated oat (Avena sativa L.) were identified using near-isogenic lines and bulked segregant analysis. Six markers for Pc48, the closest being 6 cM away, were found in a Pendek-39 × Pendek-48 (Pendek3948) population, but none was found in a Pendek-48 × Pendek-38 (Pendek4838) population. Three markers for Pc39 were found in the Pendek3948 population, one of which cosegregated with the gene. This same marker was found to be 6 cM away from the gene in an OT328 × Dumont (OT328Du) population. Nine markers for Pc38 were found in the Pendek4838 population, eight of which are within 2 cM of the gene. One other marker for Pc38 was found in the OT328Du population; however, comparative mapping suggests that the Pc38 region in OT328Du is in a different location than that in Pendek4838. A number of markers unlinked to the genes under study formed linkage groups in both the Pendek3948 and Pendek4838 populations. Four of these show homology or homoeology to each other and to the Pc39 region in Pendek3948. Two RFLP clones closely linked to Pc38 code for a putative leucine-rich repeat transmembrane protein kinase and a cre3 resistance gene analogue. This study provides information to support molecular breeding in oat, and contributes to ongoing research into genomic regions associated with fungal pathogen resistance.  相似文献   

13.
Doubled haploid (DH) progeny from a cross between the scald susceptible barley (Hordeum vulgare L.) cultivar Ingrid and the resistant accession CI 11549 (Nigrinudum) was evaluated for resistance in the pathogen Rhynchosporium secalis (Oudem) J.J. Davis. Two linked and incompletely dominant loci confer resistance CI 11549 against isolate 4004. One is an allele at the complex Rrs1 locus on chromosome 3H close to the centromere; the other is located 22 cM distally on the long arm. The latter locus is designated Rrs4. In BC3-lines into Ingrid from CI 2222 (another Nigrinudum) resistance seems governed by one locus close to the telomeric region of chromosome 7H, probably allelic to Rrs2. In neither case did we find any trace of the recessive gene rh8 reported to be present in Nigrinudum. Various resistance donors of Ethiopian origin designated as Nigrinudum, Jet or Abyssinian were identical to a great extent with respect to markers, but differed in resistance to different isolates of scald or in barley yellow dwarf virus (BYDV) resistance. The implications for their use as differentials in scald tests and screening of germplasm collections are discussed.  相似文献   

14.
Summary The inheritance of resistance to brown planthopper, Nilaparvata lugens (Stol.), of 20 rice cultivars was studied. Single dominant genes that are allelic to Bph 3 condition the resistance in cultivars Ptb 19, Gangala (Acc. 7733), Gangala (Acc. 15207), Horana Mawee, Kuruhondarwala, Mudu Kiriyal and Muthumanikam. Single recessive genes that are allelic to bph 4 govern the resistance in cultivars Gambada Samba, Heenhoranamawee, Hotel Samba, Kahata Samba, Kalukuruwee, Lekam Samba, Senawee, Sulai, Thirissa and Vellai Illankali. The resistance in Ptb 33, Sudu Hondarwala, and Sinna Sivappu is governed by one dominant and one recessive gene which segregate independently of each other. The dominant resistance genes in these cultivars appear allelic to either Bph 1 or Bph 3. Similarly, the recessive genes in these cultivars seem allelic to either bph 2 or bph 4. Further investigations are needed to conclusively determine the allelic relationships of resistance genes in Ptb 33, Sudu Hondarwala and Sinna Sivappu.  相似文献   

15.
Pepper (Capsicum chinense Jacq.) has been reported to be an important reservoir of resistance genes to tomato spotted wilt virus (TSWV). The genes for TSWV resistance present in three C. chinense lines (PI 152225, PI 159236 and Panca) were investigated for allelism. All resistant lines were crossed with each other. Parents, F1, backcrosses and F2 populations (including reciprocals) developed from those crosses were mechanically inoculated with a highly virulent TSWV isolate. Susceptible C. annuum cv Magda was used to check inoculum virulence. Fifty plants of the F1 hybrids; Magda x PI 152225, Magda x PI 159236, and Magda x 'Panca, were also inoculated with the TSWV isolate. The resistance response in all C. chinense sources was associated with a localized, hypersensitive-like reaction that was phenotypically expressed as a prompt formation of large local lesions accompanied by premature leaf abscission. All F1 generations presented a final score of resistant; indicating that the expression of resistance to TSWV is conditioned by a dominant gene regardless of the source. The absence of segregation for resistance to TSWV that was observed in all generations of the crosses between C. chinense lines indicated that either a tightly linked group of genes exists or that the resistance is governed by the same single major gene (probably the already described Tsw gene). Previous reports have indicated that the Tsw gene is not effective against tospovirus members of serogroup II, i.e. tomato chlorotic spot virus (TCSV) and groundnut ring spot virus (GRSV). In the assay described here, all of the C. chinense lines showed, after mechanical inoculation, an identical susceptibility response to the TCSV and GRSV isolates.  相似文献   

16.
Linkage of randomly amplified polymorphic DNA (RAPD) markers with a single dominant gene for resistance to black root rot (Chalara elegans Nag Raj and Kendrick; Syn. Thielaviopsis basicola [Berk. and Broome] Ferraris) of tobacco (Nicotiana tabacum L.), which was transferred from N. debneyi Domin, was investigated in this study. There were 2594 repeatable RAPD fragments generated by 441 primers on DNAs of Delgold tobacco, a BC5F8 near isogenic line (NIL) carrying the resistance gene in a Delgold background, and PB19, the donor parent of the resistance gene. Only 7 of these primers produced eight RAPD markers polymorphic between Delgold and PB19, indicating there are few RAPD polymorphisms between them despite relatively dissimilar pedigrees. Five of the eight RAPD markers were not polymorphic between Delgold and the NIL. All of these markers proved to be unlinked with the resistance gene in F2 linkage tests. Of the remaining three RAPD markers polymorphic between Delgold and the NIL, two were shown to be strongly linked with the resistance gene; one in coupling and the other in repulsion. Application of the two RAPDs in the elimination of linkage drag associated with the N. debneyi resistance gene and marker-assisted selection for the breeding of new tobacco cultivars with the resistance gene is discussed.  相似文献   

17.
Sheath blight, caused by Rhizoctonia solani, is one of the most important diseases of rice. Despite extensive searches of the rice germ plasm, the major gene(s) which give complete resistance to the fungus have not been identified. However, there is much variation in quantitatively inherited resistance to R. solani, and this type of resistance can offer adequate protection against the pathogen under field conditions. Using 255 F4 bulked populations from a cross between the susceptible variety Lemont and the resistant variety Teqing, 2 years of field disease evaluation and 113 well-distributed RFLP markers, we identified six quantitative trait loci (QTLs) contributing to resistance to R. solani. These QTLs are located on 6 of the 12 rice chromosomes and collectively explain approximately 60% of the genotypic variation or 47% of the phenotypic variation in the LemontxTeqing cross. One of these resistance QTLs (QSbr4a), which accounted for 6% of the genotypic variation in resistance to R. solani, appeared to be independent of associated morphological traits. The remaining five putative resistance loci (QSbr2a, QSbr3a, QSbr8a, QSbr9a and QSbr12a) all mapped to chromosomal regions also associated with increased plant height, three of which were also associated with QTLs causing later heading. This was consistent with the observation that heading date and plant height accounted for 47% of the genotypic variation in resistance to R. solani in this population. There were also weak associations between resistance to R. solani and leaf width, which were likely due to linkage with a QTL for this trait rather than to a physiological relationship.  相似文献   

18.
Powdery mildew poses a serious problem for apple growers, and resistance to the disease is a major objective in breeding programmes for cultivar improvement. As selective pressure allows pathogens to overcome previously reliable resistances, there is a need for the introduction of novel resistance genes into new breeding lines. This investigation is concerned with the identification of the first set of molecular markers linked to the gene for mildew resistance, Pl-d, from the accession D12. As no prior information on the map position or markers for Pl-d were available, a bulked-segregant approach was used to test 49 microsatellite primers, 176 amplified fragment length polymorphism (AFLP) primers and 80 random amplified polymorphic DNA (RAPD) primers in a progeny segregating for Pl-d resistance, Fiesta (susceptible) × A871-14 (Worcester Pearmain × D12). The segregations of the markers identified in the resistant and susceptible bulks were scored in the progeny, then the recombination fractions between Pl-d and the most tightly linked markers were calculated and a map prepared. Three AFLP, one RAPD and two microsatellite markers were identified. One AFLP was developed into a sequence-characterised amplified region marker, while the microsatellites CH03C02 and CH01D03 were flanking markers, 7 and 11 recombination units, respectively, from Pl-d. Two more distant microsatellites on the same linkage group, CH01D09 and CH01G12, confirmed the orientation of the markers on the linkage group. These microsatellites place Pl-d on the bottom of linkage group 12 in published apple maps, a region where a number of other disease resistance genes have been identified.  相似文献   

19.
Summary Clear-plaque phage c, attacking bacitracin-producing strains of B. licheniformis, yields spontaneous temperate mutants at high frequency; the temperate mutants fall into several classes phenotypically different in plaque morphology and properties of lysogenised bacteria. The most common phenotype 3 has DNA restriction fragment patterns identical with those of the parent c; some less common temperate forms, i.e. 1 and 2, produce different restriction fragment patterns, sugesting that a part of the original c DNA has been reorganized or replaced by some foreign genetic material. The changed fragment pattern remains stable upon subsequent passaging of the phage or of the lysogenic bacteria. Neither class of temperate phage mutants gives clearplaque revertants at measurable frequency. Lysogenisation of bacteria with any class of temperate phage confers immunity to all temperate forms and to c; virulent mutants vir, which plate with 100% efficiency on lysogens for 1 and 2 but not for 3, occur in stocks of c at a frequency of 10–7. The mutation from c to vir is not accompanied by any change of the restriction fragment patterns of DNA.  相似文献   

20.
Outside-out configuration of the patch clamp technique was used to test whether an intracellular application of G protein activator (GTPS) affects ATP-activated Ca2+-permeable channels in rat macrophages without any agonist in the bath solution. With 145 mm K+ (pCa 8.0) in the pipette solution, activity of channels permeable to a variety of divalent cations and Na+ was observed and general channel characteristics were found to be identical to those of ATP-activated ones. Absence of extracellular ATP makes it possible to avoid the influence of ATP receptor desensitization and to study the channel selectivity using a number of divalent cations (105 mm) and Na+ (145 mm) as the charge carriers. Permeability sequence estimated by extrapolated reversal potential measurements was: Ca2+ Ba2+ Mn2+ Sr2+ Na+ K+ = 68 30 26 10 3.5 1. Slope conductances (in pS) for permeant ions rank as follows: Ca2+ Sr2+ Na+ Mn2+ Ba2+ = 19 18 14 12 10. Unitary Ca2+ currents display a tendency to saturate with the Ca2+ concentration increase with apparent dissociation constant (K d ) of 10 mm. No block of Na+ permeation by extracellular Ca2+ in millimolar range was found. The data obtained suggest that (i) activation of some G protein is sufficient to gate the channels without the ATP receptor being occupied, (ii) the ATP receptor activation results in the gating of a special channel with the properties that differ markedly from those of the receptoroperated or voltage-gated Ca2+-permeable channels on the other cell types.DeceasedThe authors are grateful to K. Kiselyov and A. Mamin for technical assistance. The work was supported by the Russian Basic Research Foundation, Grant N 93-04-21722 and was made possible in part by Grant N R4A000 from the International Science Foundation.  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