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.
Vertebrate akirin genes usually form a family with one-to-three members that regulate gene expression during the innate immune response, carcinogenesis and myogenesis. We recently established that an expanded family of eight akirin genes is conserved across salmonid fish. Here, we measured mRNA levels of the akirin family of Atlantic salmon (Salmo salar L.) during the differentiation of primary myoblasts cultured from fast-skeletal muscle. Using hierarchical clustering and correlation, the data was positioned into a network of expression profiles including twenty further genes that regulate myogenesis. akirin1(2b) was not significantly regulated during the maturation of the cell culture. akirin2(1a) and 2(1b), along with IGF-II and several igfbps, were most highly expressed in mononuclear cells, then significantly and constitutively downregulated as differentiation proceeded and myotubes formed/matured. Conversely, akirin1(1a), 1(1b), 1(2a), 2(2a) and 2(2b) were expressed at lowest levels when mononuclear cells dominated the culture and highest levels when confluent layers of myotubes were evident. However, akirin1(2a) and 2(2a) were first upregulated earlier than akirin1(1a), 1(1b) and 2(2b), when rates of myoblast proliferation were highest. Interestingly, akirin1(1b), 1(2a), 2(2a) and 2(2b) formed part of a module of co-expressed genes involved in muscle differentiation, including myod1a, myog, mef2a, 14-3-3β and 14-3-3γ. All akirin paralogues were expressed ubiquitously across ten tissues, although mRNA levels were regulated between cell-types and family members. Gene expression patterns were often highly correlated between akirin paralogues, suggesting that natural selection has maintained an intricate network of co-regulation among family members. We concluded that the Atlantic salmon akirin family performs a multifaceted role during myogenesis and has physiological functions spanning many cell-types.  相似文献   

2.
The O-glycosidically-linked carbohydrate units of glycophorin from bovine erythrocyte membrane were released by alkaline borohydride treatment. These oligosaccharides were separated into the neutral fractions and the acidic fractions by ion-exchange chromatography followed by gel filtration. The two acidic fractions (fractions 10 and 13) which have the smallest molecular weight in acidic oligosaccharides, were further purified by gel filtration on Bio-Gel P-4 column. Two acidic oligosaccharides (fractions 10-I and 10-II), heptasaccharides, were separated by gel filtration on a Bio-Gel P-4 column from fraction 10. These structures were determined by methylation analyses, nitrous acid deamination after hydrazinolysis and Smith degradation after desialylation. In addition, the structures were also analyzed by direct-probe mass spectrometry of the permethylated derivatives before and after desialylation. These studies indicated that one of them (fraction 10-I) was NeuNGcα(2→3)Galβ(1→4)GlcNAcβ(1→3)Galβ(1→4)GlcNAcβ(1→3)Galβ(1→3) GalNAcol and another heptasaccharide (fraction 10-II) was Galβ(1→4)GlcNAcβ(1→3)Galβ(1→3) [NeuNGcα(2→3)Galβ(1→4)GlcNAcβ(1→6)]GalNAcol. Athough another acidic fraction (fraction 13) was obtained as a single peak on a Bio-Gel P-4 column, it appeared to be the mixture of a heptasaccharide, NeuNGcα(2→3)Galβ(1→4)GlcNAcβ(1→3 or 6)[Galβ(1→4)GlcNAcβ(1→6 or 3)]Galβ(1→3)GalNAcol and an oligosaccharide similar to fraction 10-II, by analysis of two products obtained by Smith degradation after desialylation.  相似文献   

3.
Twenty-four male internals(locus of control) and 24 externals were instructed to increase finger temperature under one of three conditions:(1) contingent feedback(CF),(2) yoked sham feedback(YF), or(3) no feedback(NF). Five 13-min training sessions were given. Feedback was then removed and subjects tested for voluntary temperature control. Finally, the cold pressor test, a laboratory analogue of natural cold stress, was administered under no-feedback conditions. Results demonstrated voluntary control of peripheral temperature following contingent feedback training, but not after yoked feedback temperature training. Contrary to expectation, the acquisition of voluntary control did not attenuate the stress response to thermal pain. Differences between internals and externals throughout the study were generally insignificant.  相似文献   

4.
Reaction of Y5O(OiPr)13 (“yttrium iso-propoxide”) with one molar equivalent of isopropyl acetoacetate (iprac) per Y resulted in the formation of Y9O(OH)9(OiPr)8(iprac)8, a rare example of an yttrium alkoxo/hydroxo/oxo cluster. Reaction in a 1:3 molar ratio gave Y4(OH)2(iprac)10 and Y6(OH)6(iprac)12 instead. A fourth cluster, Y9O(OH)9(iprac)16, structurally closely related to Y9O(OH)9(OiPr)8(iprac)8, was obtained upon recrystallization of Y4(OH)2(iprac)10 from CDCl3.  相似文献   

5.
Cross-Correlation Functions for a Neuronal Model   总被引:5,自引:1,他引:4       下载免费PDF全文
Cross-correlation functions, RXY(t,τ), are obtained for a neuron model which is characterized by constant threshold θ, by resetting to resting level after an output, and by membrane potential U(t) which results from linear summation of excitatory postsynaptic potentials h(t). The results show that: (1) Near time lag τ = 0, RXY(t,τ) = fU [θ-h(τ), t + τ] {h′(τ) + EU [u′(t + τ)]} for positive values of this quantity, where fU(u,t) is the probability density function of U(t) and EU [u′(t + τ)] is the mean value function of U′(t + τ). (2) Minima may appear in RXY(t,τ) for a neuron subjected only to excitation. (3) For large τ, RXY(t,τ) is given approximately by the convolution of the input autocorrelation function with the functional of point (1). (4) RXY(t,τ) is a biased estimator of the shape of h(t), generally over-estimating both its time to peak and its rise time.  相似文献   

6.
3种海拔高度茶园中2种害虫与其天敌间的数量和空间关系   总被引:4,自引:0,他引:4  
用灰色系统分析法和生态位分析法对安徽省潜山县高、中、低3种海拔茶园中2种主要害虫与其天敌在数量和空间上的相互关系进行研讨,经综合排序得出:低海拔茶园假眼小绿叶蝉的主要天敌是锥腹肖蛸、茶色新圆蛛和鞍型花蟹蛛,中海拔茶园为八点球腹蛛、锥腹肖蛸和鞍型花蟹蛛,高海拔茶园为锥腹肖蛸、茶色新圆蛛和草间小黒蛛。低海拔茶园柑橘粉虱的主要天敌是八点球腹蛛、茶色新圆蛛和异色瓢虫,中海拔茶园为鞍型花蟹蛛、八点球腹蛛和茶色新圆蛛,高海拔茶园为锥腹肖蛸、八点球腹蛛和斑管巢蛛。上述几种天敌与2种主要害虫在数量和空间上有显著的追随关系。不同海拔高度茶园中2种主要害虫的主要天敌种类存在差异。  相似文献   

7.
Based on a number of experiments it is concluded that the fluorescein labeled β-heptapeptide fluoresceinyl-NH-CS-(S)-β3hAla-(S)-β3hArg-(R)-β3hLeu-(S)-β3hPhe-(S)-β3hAla-(S)-β3hAla-(S)-β3hLys-OH translocates across lipid vesicle bilayers formed from DOPC (1,2-dioleoyl-sn-glycero-3-phosphocholine). The conclusion is based on the following observations: (i) addition of the peptide to the vicinity of micrometer-sized giant vesicles leads to an accumulation of the peptide inside the vesicles; (ii) if the peptide is injected inside individual giant vesicles, it is released from the vesicles in a time dependent manner; (iii) if the peptide is encapsulated within sub-micrometer-sized large unilamellar vesicles, it is released from the vesicles as a function of time; (iv) if the peptide is submitted to immobilized liposome chromatography, the peptide is retained by the immobilized DOPC vesicles. Furthermore, the addition of the peptide to calcein-containing DOPC vesicles does not lead to significant calcein leakage and vesicle fusion is not observed. The finding that derivatives of the β-heptapeptide (S)-β3hAla-(S)-β3hArg-(R)-β3hLeu-(S)-β3hPhe-(S)-β3hAla-(S)-β3hAla-(S)-β3hLys-OH can translocate across phospholipid bilayers is supported by independent measurements using Tb3+-containing large unilamellar vesicles prepared from egg phosphatidylcholine and wheat germ phosphatidylinositol (molar ratio of 9:1) and a corresponding peptide that is labeled with dipicolinic acid instead of fluorescein. The experiments show that this dipicolinic acid labeled β-heptapeptide derivative also permeates across phospholipid bilayers. The possible mechanism of the translocation of the particular β-heptapeptide derivatives across the membrane of phospholipid vesicles is discussed within the frame of the current understanding of the permeation of certain oligopeptides across simple phospholipid bilayers.  相似文献   

8.
The ref(2)P gene of Drosophila melanogaster is implicated in sigma rhabdovirus multiplication. Two common alleles of ref(2)P are known, ref(2)P 0 which permits sigma virus multiplication and ref(2)P pwhich is restrictive for most sigma virus strains. This gene maps to the cytogenetic region 37E3-F3. Using Df(2L)E55 (=Df(2L)37D2-El;37F5-38A1), we have screened for lethal, semi-lethal and visible mutations following diepoxybutane (DEB) or ethyl methanesulfonate (EMS) mutagenesis. Our data confirm than DEB is mor efficient than EMS at inducing deletions. The mutations obtained in this region define 14 complementation groups. One of them, l(2)37Dh, appears to be a general enhancer of Minute and Minute-like mutations. None of the mutations were allelic to the ref(2)P locus. Loss-of-function alleles of ref(2)P (called null) were selected following DEB mutagenesis. Homozygous or hemizygous ref(2)P nullflies are male sterile. These flies, like homozygous or hemizygous ref(2)P 0flies, are fully permissive for sigma virus replication. We suggest that the ref(2)P products interact with viral products, but that this interaction is not necessary for an efficient viral cycle.  相似文献   

9.
A total of 30 Megasphaera elsdenii strains, selectively isolated from the feces of organically raised swine by using Me109 M medium, and one bovine strain were analyzed for tetracycline resistance genotypic and phenotypic traits. Tetracycline-resistant strains carried tet(O), tet(W), or a tet gene mosaic of tet(O) and tet(W). M. elsdenii strains carrying tet(OWO) genes exhibited the highest tetracycline MICs (128 to >256 μg/ml), suggesting that tet(O)-tet(W) mosaic genes provide the selective advantage of greater tetracycline resistance for this species. Seven tet genotypes are now known for M. elsdenii, an archetype commensal anaerobe and model for tet gene evolution in the mammalian intestinal tract.  相似文献   

10.
Multidrug-resistant enterococci are considered crucial drivers for the dissemination of antimicrobial resistance determinants within and beyond a genus. These organisms may pass numerous resistance determinants to other harmful pathogens, whose multiple resistances would cause adverse consequences. Therefore, an understanding of the coexistence epidemiology of resistance genes is critical, but such information remains limited. In this study, our first objective was to determine the prevalence of principal resistance phenotypes and genes among Enterococcus faecalis isolated from retail chicken domestic products collected throughout Japan. Subsequent analysis of these data by using an additive Bayesian network (ABN) model revealed the co-appearance patterns of resistance genes and identified the associations between resistance genes and phenotypes. The common phenotypes observed among E. faecalis isolated from the domestic products were the resistances to oxytetracycline (58.4%), dihydrostreptomycin (50.4%), and erythromycin (37.2%), and the gene tet(L) was detected in 46.0% of the isolates. The ABN model identified statistically significant associations between tet(L) and erm(B), tet(L) and ant(6)-Ia, ant(6)-Ia and aph(3’)-IIIa, and aph(3’)-IIIa and erm(B), which indicated that a multiple-resistance profile of tetracycline, erythromycin, streptomycin, and kanamycin is systematic rather than random. Conversely, the presence of tet(O) was only negatively associated with that of erm(B) and tet(M), which suggested that in the presence of tet(O), the aforementioned multiple resistance is unlikely to be observed. Such heterogeneity in linkages among genes that confer the same phenotypic resistance highlights the importance of incorporating genetic information when investigating the risk factors for the spread of resistance. The epidemiological factors that underlie the persistence of systematic multiple-resistance patterns warrant further investigations with appropriate adjustments for ecological and bacteriological factors.  相似文献   

11.
The synthesis of tetrapeptide-based β-turn mimetics containing spirocyclic glucose-templated 3-hydroxyproline hybrids Glc3′(S)-5′(R)(CH2OH)HypH and Glc3′(S)-5′(S)(CH2OH)HypH as proline mimetics is presented. NMR-based conformational analysis of Ac-Leu-d-Phe-[Glc3′(S)-5′(R)(CH2OH)HypH]-Val-NMe2 and Ac-Leu-d-Phe-[Glc3′(S)-5′(S)(CH2OH)HypH]-Val-NMe2 demonstrates the presence of β-turn conformations. Different turn structures were observed by changing the stereochemistry at 5′-position of Glc3′(S)-5′(R)(CH2OH)HypH. The major prolyl amide cis isomer of glucose-protected tetrapeptide Ac-Leu-d-Phe-[Glc(MOM)43′(S)-5′(R)(CH2OMOM)HypH]-Val-NMe211 and glucose unprotected Ac-Leu-d-Phe-[Glc3′(S)-5′(R)(CH2OH)HypH]-Val-NMe213 forms a type VI β-turn conformation. In contrast, the major prolyl amide trans rotamer of tetrapeptide Ac-Leu-d-Phe-[Glc(MOM)43′(S)-5′(S)(CH2OMOM)HypH]-Val-NMe212 conserves a similar β-turn conformation as the Gramicidin S-based peptide fragment Ac-Leu-d-Phe-Pro-Val-NMe216.  相似文献   

12.
Germination of Bacillus spores requires degradation of a modified layer of peptidoglycan (PG) termed the spore cortex by two redundant cortex-lytic enzymes (CLEs), CwlJ and SleB, plus SleB''s partner protein, YpeB. In this study, in vitro and in vivo analyses have been used to clarify the roles of individual SleB and YpeB domains in PG degradation. Purified mature Bacillus cereus SleB without its signal sequence (SleBM) and the SleB C-terminal catalytic domain (SleBC) efficiently triggered germination of decoated Bacillus megaterium and Bacillus subtilis spores lacking endogenous CLEs; previously, SleB''s N-terminal domain (SleBN) was shown to bind PG but have no enzymatic activity. YpeB lacking its putative membrane anchoring sequence (YpeBM) or its N- and C-terminal domains (YpeBN and YpeBC) alone did not exhibit degradative activity, but YpeBN inhibited SleBM and SleBC activity in vitro. The severe germination defect of B. subtilis cwlJ sleB or cwlJ sleB ypeB spores was complemented by ectopic expression of full-length sleB [sleB(FL)] and ypeB [ypeB(FL)], but normal levels of SleBFL in spores required normal spore levels of YpeBFL and vice versa. sleB(FL) or ypeB(FL) alone, sleB(FL) plus ypeB(C) or ypeB(N), and sleB(C) or sleB(N) plus ypeB(FL) did not complement the cortex degradation defect in cwlJ sleB ypeB spores. In addition, ectopic expression of sleB(FL) or cwlJ(FL) with a Glu-to-Gln mutation in a predicted active-site residue failed to restore the germination of cwlJ sleB spores, supporting the role of this invariant glutamate as the key catalytic residue in SleB and CwlJ.  相似文献   

13.
Resistances to tetracycline and mercury were identified in an environmental strain of Serratia marcescens isolated from a stream highly contaminated with heavy metals. As a step toward addressing the mechanisms of coselection of heavy metal and antibiotic resistances, the tetracycline resistance determinant was cloned in Escherichia coli. Within the cloned 13-kb segment, the tetracycline resistance locus was localized by deletion analysis and transposon mutagenesis. DNA sequence analysis of an 8.0-kb region revealed a novel gene [tetA(41)] that was predicted to encode a tetracycline efflux pump. Phylogenetic analysis showed that the TetA(41) protein was most closely related to the Tet(39) efflux protein of Acinetobacter spp. yet had less than 80% amino acid identity with known tetracycline efflux pumps. Adjacent to the tetA(41) gene was a divergently transcribed gene [tetR(41)] predicted to encode a tetracycline-responsive repressor protein. The tetA(41)-tetR(41) intergenic region contained putative operators for TetR(41) binding. The tetA(41) and tetR(41) promoters were analyzed using lacZ fusions, which showed that the expression of both the tetA(41) and tetR(41) genes exhibited TetR(41)-dependent regulation by subinhibitory concentrations of tetracycline. The apparent lack of plasmids in this S. marcescens strain, as well as the presence of metabolic genes adjacent to the tetracycline resistance locus, suggested that the genes were located on the S. marcescens chromosome and may have been acquired by transduction. The cloned Tet 41 determinant did not confer mercury resistance to E. coli, confirming that Tet 41 is a tetracycline-specific efflux pump rather than a multidrug transporter.  相似文献   

14.
Two routes to protected derivatives of 2,5-anhydroallitol were investigated. The first route, involving a two-step reduction of 2,5-anhydro-6-O-benzoyl-3,4-O-isopropylidene-D-allonitrile (4), gave a mixture of 2,5-anhydro-6-O-benzoyl-3,4-O-isopropylidene-D-altritol (7) and a lesser amount of the desired 2,5-anhydro-6-O-benzoyl-3,4-O-isopropylidene-D-allitol (6). Isomerization was shown to occur in the first reduction step—treatment of the nitrile 4 with Raney nickel, sodium hypophosphite, and acetic acid. The second route gave isomerically pure 2,5-anhydro-3,4,6-tri-O-benzyl-D-allitol (21) via reduction of the corresponding ethyl allonate (18).  相似文献   

15.
The glucolipids formed upon feeding (U-14C)glucose to embryos of Zea mays were partially characterized with respect to: (a) metabolic turnover, (b) acid lability, (c) phosphorus content, (d) chromatographic properties, and (e) hydrolysis products. The chloroform-methanol-soluble assimilated radioactivity was examined specifically for occurrence of a glycosylated prenol phosphate. With the extraction conditions used, no evidence was found for formation of a glucosylated prenol phosphate. Several, as yet unidentified, acid-labile glucolipids undergoing metabolic turnover were observed. Four diglycerides were characterized as hydrolysis products of a fraction that contained 14C-glucose and phosphorus, and was subject to metabolic turnover. Examination of the 1-butanol-soluble glucolipids from pea (Pisum sativum) seedlings also demonstrated anionic glucolipids, evidencing metabolic turnover but none with the properties of glucosylated prenol phosphate.  相似文献   

16.
Quantitative real-time PCR (qRT-PCR) is a powerful technique to quantify gene expression. To standardize gene expression studies and obtain more accurate qRT-PCR analysis, normalization relative to consistently expressed housekeeping genes (HKGs) is required. In this study, ten candidate HKGs including elongation factor 1 α (EF1A), ribosomal protein L11 (RPL11), ribosomal protein L14 (RPL14), ribosomal protein S8 (RPS8), ribosomal protein S23 (RPS23), NADH-ubiquinone oxidoreductase (NADH), vacuolar-type H+-ATPase (ATPase), heat shock protein 70 (HSP70), 18S ribosomal RNA (18S), and 12S ribosomal RNA (12S) from the cowpea aphid, Aphis craccivora Koch were selected. Four algorithms, geNorm, Normfinder, BestKeeper, and the ΔCt method were employed to evaluate the expression profiles of these HKGs as endogenous controls across different developmental stages and temperature regimes. Based on RefFinder, which integrates all four analytical algorithms to compare and rank the candidate HKGs, RPS8, RPL14, and RPL11 were the three most stable HKGs across different developmental stages and temperature conditions. This study is the first step to establish a standardized qRT-PCR analysis in A. craccivora following the MIQE guideline. Results from this study lay a foundation for the genomics and functional genomics research in this sap-sucking insect pest with substantial economic impact.  相似文献   

17.
18.

Background

Protozoan parasites of the genus Leishmania cause a large spectrum of clinical manifestations known as Leishmaniases. These diseases are increasingly important public health problems in many countries both within and outside endemic regions. Thus, an accurate differential diagnosis is extremely relevant for understanding epidemiological profiles and for the administration of the best therapeutic protocol.

Methods/Principal Findings

Exploring the High Resolution Melting (HRM) dissociation profiles of two amplicons using real time polymerase chain reaction (real-time PCR) targeting heat-shock protein 70 coding gene (hsp70) revealed differences that allowed the discrimination of genomic DNA samples of eight Leishmania species found in the Americas, including Leishmania (Leishmania) infantum chagasi, L. (L.) amazonensis, L. (L.) mexicana, L. (Viannia) lainsoni, L. (V.) braziliensis, L. (V.) guyanensis, L. (V.) naiffi and L. (V.) shawi, and three species found in Eurasia and Africa, including L. (L.) tropica, L. (L.) donovani and L. (L.) major. In addition, we tested DNA samples obtained from standard promastigote culture, naturally infected phlebotomines, experimentally infected mice and clinical human samples to validate the proposed protocol.

Conclusions/Significance

HRM analysis of hsp70 amplicons is a fast and robust strategy that allowed for the detection and discrimination of all Leishmania species responsible for the Leishmaniases in Brazil and Eurasia/Africa with high sensitivity and accuracy. This method could detect less than one parasite per reaction, even in the presence of host DNA.  相似文献   

19.
M. Ashburner  S. Tsubota    R. C. Woodruff 《Genetics》1982,102(3):401-420
Exchange mapping locates the dominant mutation Scutoid to the right of Adh on chromosome arm 2L of D. melanogaster. However, deletion mapping indicates that Sco is to the left of Adh. The phenotype of Sco is sensitive to mutation, or deletion, of noc+ and of three genes, el, l(2)br22, and l(2)br29 mapping immediately distal to noc. The four contiguous loci, el, l(2)br22, l(2)br29 and noc, although separable by deletion end points, interact, because certain (or all) alleles of these four loci show partial failure of complementation, or even negative complementation. The simplest hypothesis is that Sco is a small reciprocal transposition, the genes noc, osp, and Adh exchanging places with three genes normally mapping proximal to them: l(2)br34, l(2)br35 and rd. The Sco phenotype is thought to result from a position effect at the newly created noc/l(2)br28 junction.  相似文献   

20.
In spite of a global concern on the transfer of antibiotic resistances (AR) via the food chain, limited information exists on this issue in species of Leuconostoc and Weissella, adjunct cultures used as aroma producers in fermented foods. In this work, the minimum inhibitory concentration was determined for 16 antibiotics in 34 strains of dairy origin, belonging to Leuconostoc mesenteroides (18), Leuconostoc citreum (11), Leuconostoc lactis (2), Weissella hellenica (2), and Leuconostoc carnosum (1). Atypical resistances were found for kanamycin (17 strains), tetracycline and chloramphenicol (two strains each), and erythromycin, clindamycin, virginiamycin, ciprofloxacin, and rifampicin (one strain each). Surprisingly, L. mesenteroides subsp. mesenteroides LbE16, showed resistance to four antibiotics, kanamycin, streptomycin, tetracycline and virginiamycin. PCR analysis identified tet(S) as responsible for tetracycline resistance in LbE16, but no gene was detected in a second tetracycline-resistant strain, L. mesenteroides subsp. cremoris LbT16. In Leuconostoc mesenteroides subsp. dextranicum LbE15, erythromycin and clindamycin resistant, an erm(B) gene was amplified. Hybridization experiments proved erm(B) and tet(S) to be associated to a plasmid of ≈35 kbp and to the chromosome of LbE15 and LbE16, respectively. The complete genome sequence of LbE15 and LbE16 was used to get further insights on the makeup and genetic organization of AR genes. Genome analysis confirmed the presence and location of erm(B) and tet(S), but genes providing tetracycline resistance in LbT16 were again not identified. In the genome of the multi-resistant strain LbE16, genes that might be involved in aminoglycoside (aadE, aphA-3, sat4) and virginiamycin [vat(E)] resistance were further found. The erm(B) gene but not tet(S) was transferred from Leuconostoc to Enterococcus faecalis both under laboratory conditions and in cheese. This study contributes to the characterization of AR in the Leuconostoc-Weissella group, provides evidence of the genetic basis of atypical resistances, and demonstrates the inter-species transfer of erythromycin resistance.  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