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.
Wilma K. Olson 《Biopolymers》1975,14(9):1797-1810
The dependence of the unperturbed dimensions of randomly coiling polynucleotides on the rotations about the phosphodiester linkages of the chain has been examined in order to understand the conformational discrepancies, set forth in paper I, regarding these angles (ω′ and ω). Large values of the characteristic ratio 〈r20/nl2 , which agree with the experimental behavior of the chain, are obtained only if a sizeable proportion of the polymer residues have trans ω′ values. The asymmetric torsional potential that is believed to arise from gauche effects associated with the P-O bonds has been approximated using a hard core model. The calculated characteristic ratio exhibits a strong dependence upon the magnitude of this torsional barrier (separating trans and gauche conformations) and shows agreement with experimental values for polyribonucleotides only if this energy difference is 1 kcal/mol or less.  相似文献   

2.
Abstract: The voltage-dependent calcium channels present in mammalian and chicken brain synaptosomes were characterized pharmacologically using specific blockers of L-type channels (1,4-dihydropyridines), N-type channels (ω-conotoxin GVIA), and P-type channels [funnel web toxin (FTX) and ω-agatoxin IVA]. K+-induced Ca2+ uptake by chicken synaptosomes was blocked by ω-conotoxin GVIA (IC50 = 250 nM). This toxin at 5 µM did not block Ca2+ entry into rat frontal cortex synaptosomes. FTX and ω-agatoxin IVA blocked Ca2+ uptake by rat synaptosomes (IC50 = 0.17 µl/ml and 40 nM, respectively). Likewise, in chicken synaptosomes, FTX and ω-agatoxin IVA affected Ca2+ uptake. FTX (3 µl/ml) exerted a maximal inhibition of 40% with an IC50 similar to the one obtained in rat preparations, whereas with ω-agatoxin IVA saturation was not reached even at 5 µM. In chicken preparations, the combined effect of saturating concentrations of FTX (1 µl/ml) and different concentrations of ω-conotoxin GVIA showed no additive effects. However, the effect of saturating concentrations of FTX and ω-conotoxin GVIA was never greater than the one observed with ω-conotoxin GVIA. We also found that 60% of the Ca2+ uptake by rat and chicken synaptosomes was inhibited by ω-conotoxin MVIID (1 µM), a toxin that has a high index of discrimination against N-type channels. Conversely, nitrendipine (10 µM) had no significant effect on Ca2+ uptake in either the rat or the chicken. In conclusion, Ca2+ uptake by rat synaptosomes is potently inhibited by different P-type Ca2+ channel blockers, thus indicating that P-type channels are predominant in this preparation. In contrast, Ca2+ uptake by chicken synaptosomes is sensitive to ω-conotoxin GVIA, FTX, ω-agatoxin IVA, and ω-conotoxin MVIID. This suggests that a channel subtype with a mixed pharmacology is present in chicken synaptosomes.  相似文献   

3.
Bai  Xin-Feng  Lv  Xue-Lian  Liu  Xue  Cui  Ting-Ting  Zhang  Mian-Song  Ding  Ning  Liu  Chang-Heng  Jia  Ai-Rong 《Antonie van Leeuwenhoek》2022,115(4):497-503

A Gram-stain negative, aerobic, rod-shaped bacterium, designated 126T, was isolated from the intestinal content of a sea cucumber, Apostichopus japonicus, in China. Strain 126T was found to grow optimally at 25–28 °C and pH 7.5–8.0 in marine 2216 E medium, with tolerance of 1–7% (w/v) NaCl. Strain 126T is motile by means of one to several polar flagella. The dominant fatty acids of strain 126T were identified as C16:1 ω7c/C16:1 ω6c (29.5%), C18:1 ω7c/C18:1 ω6c (19.8%) and C16:0 (16.7%). The respiratory quinone was found to be Q-8. The polar lipid profile was found to be mainly composed of phosphatidylglycerol and phosphatidylethanolamine. The total length of the draft genome is approximately 4.2?×?106 bp, encoding 3655 genes and 3576 coding sequences. The G?+?C content of the genomic DNA is 48.0%. Phylogenetic analysis based on 16S rRNA gene sequences indicated that strain 126T belongs to the genus Neiella and is closely related to Neiella marina J221T (96.5%). Genomic comparisons of 126T to N. marina J221T revealed that they had similar genome size, G?+?C content and complement of clusters of orthologous groups. However, average nucleotide identity and digital DNA–DNA hybridization values between strains126T and N. marina J221T was 75.5% and 19.7%, which could distinguish the strains. On the basis of these phenotypic and genotypic data, strain 126T is concluded to represent a novel species, for which the name Neiella holothuriorum sp. nov. is proposed. The type strain is 126T (=?GDMCC 1.2530T?=?KCTC 82829T).

  相似文献   

4.
Wang  Guanghua  Liu  Jianfeng  Li  Yuanjin  Li  Jin  Luo  Jixin  Chen  Biao  Liao  Zhiheng  Su  Hongfei  Liang  Jiayuan  Yu  Kefu 《Antonie van Leeuwenhoek》2022,115(7):933-941

Thermal stress is considered one of the main causes of mass scleractinian coral degradation; however, it is still unknown how corals can adapt to future global warming. In this study, 11 strains of coral-associated Flavobacteria were shown to produce zeaxanthin, a carotenoid antioxidant, which may help coral holobionts to alleviate thermal stress. In addition, a novel zeaxanthin-producing Flavobacterium, designated R38T, was identified using polyphasic taxonomy. Although strain R38T shared a maximum 16S rRNA gene sequence similarity of 93% with Mesoflavibacter aestuarii KYW614T, phylogenetic analyses based on whole genome and 16S rRNA gene sequences revealed that strain R38T forms a distinct branch in a robust cluster composed of strain R38T and Leptobacterium flavescens KCTC 22160T under the family Flavobacteriaceae. Strain R38T exhibited average nucleotide identities of 70.2% and 72.5% for M. aestuarii KYW614T and L. flavescens KCTC 22160T, respectively. The only detected respiratory quinone was menaquinone 6 (MK-6). The genomic DNA G?+?C content was 33.2 mol%. The major polar lipids were phosphatidylmethylethanolamine, phosphatidylethanolamine, one unidentified ninhydrin phospholipid, three unidentified ninhydrin-positive lipids, and three unidentified lipids. The major cellular fatty acids were iso???C15:?0, iso???C15:?0 ω6c, C16:2 DMA, and C13:1 ω3c. The distinct biochemical, chemotaxonomic, phylogenetic, and phylogenomic differences from validly published taxa suggest that strain R38T represents a new species of a new genus, for which Prasinibacter corallicola gen. nov., sp. nov. is proposed. The type strain R38T (=?MCCC 1K03889T?=?KCTC 72444T).

  相似文献   

5.
The dispersion properties of a magnetoactive plasma-filled waveguide in the vicinity of the upper hybrid frequency ω1 = (ω p 2 + ω B 2 )1/21 ≤ ω < ω1 + δ) are studied. In this frequency range, the eigenfrequency spectrum and the structure of the waveguide eigenfields are difficult to analyze, because one of the transverse wavenumbers tends to infinity as the upper hybrid frequency is approached. It is shown that this problem can be solved by transforming the dispersion relation into a form independent of this transverse wavenumber. It is also found that the upper hybrid frequency itself can also be a solution to the dispersion relation. The influence of the properties of the magnetoactive plasma-filled waveguide on such solutions is studied. The structure of the fields for modes with frequencies equal to the upper hybrid frequency and an infinitely large absolute value of one of the transverse wavenumbers is analyzed.  相似文献   

6.
To obtain general rules of peptide design using α,β-dehydro-residues, a sequence with two consecutive ΔPhe-residues, Boc-L -Val-ΔPhe–ΔPhe- L -Ala-OCH3, was synthesized by azlactone method in solution phase. The peptide was crystallized from its solution in an acetone/water mixture (70:30) in space group P61 with a=b=14.912(3) Å, c= 25.548(5) Å, V=4912.0(6) Å3. The structure was determined by direct methods and refined by a full matrix least-squares procedure to an R value of 0.079 for 2891 observed [I?3σ(I)] reflections. The backbone torsion angles ?1=?54(1)°, ψ1= 129(1)°, ω1=?177(1)°, ?2 =57(1)°, ψ2=15(1)°, ω2 =?170(1)°, ?3=80(1)°, ψ3 =7(2)°, ω3=?177(1)°, ?4 =?108(1)° and ψT4=?34 (1)° suggest that the peptide adopts a folded conformation with two overlapping β-turns of types II and III′. These turns are stabilized by two intramolecular hydrogen bonds between the CO of the Boc group and the NH of ΔPhe3 and the CO of Val1 and the NH of Ala4. The torsion angles of ΔPhe2 and ΔPhe3 side chains are similar and indicate that the two ΔPhe residues are essentially planar. The folded molecules form head-to- tail intermolecular hydrogen bonds giving rise to continuous helical columns which run parallel to the c-axis. This structure established the formation of two β-turns of types II and III′ respectively for sequences containing two consecutive ΔPhe residues at (i+2) and (i+3) positions with a branched β-carbon residue at one end of the tetrapeptide.  相似文献   

7.
R Tewari  R K Nanda  G Govil 《Biopolymers》1974,13(10):2015-2035
The probability distributions of the torsional angles (Φ′, ω′, ω, Φ, and ψ), which fix the structure of nucleotide backbone, have been calculated using the results of energy calculations based on extended Huckel theory (EHT), complete neglect of differential overlap (CNDO), perturbative configuration interaction using localized orbitals (PCILO), and classical potential functions (CPF) methods. Statistical average values of the vicinal 1H? 1H, 1H? 31P, and 13C? 31P nmr coupling constants 〈J〉 have been calculated from the generalized Karplus relations using the probability distribution in the Φ′, Φ, and ψ space. Experimental 〈J〉 values for polyribouridylic acid (polyU) support the theoretical predictions for these torsional angles. Using Monte Carlo technique, random coils of single-stranded polynucleotides have been simulated and the mean-square end-to-end distance 〈r2〉 has been calculated. Molecular orbital methods (EHT, CNDO, and PCILO) suggest considerable flexibility around O? P bonds, leading to fairly small values for the characteristic ratio (C ~ 4). Observed values of the unperturbed characteristic ratio for polynucleotides are quite large (C ~ 18) suggesting a relatively rigid nucleotide backbone. The results based on molecular orbital calculations can be reconciled with the experimental values by introducing an additional stabilization of ~2 kcal mol?1 for the predicted minimum energy ragion (Φ′ ~ 240°, ω′ ~ 290°, ω 290°, Φ 180°, and ψ 60°). Such a stabilization may arise from the association of water molecules and metal ions with the phosphate group and (or) Coulomb interaction between neighboring phosphate groups. The calculations provide a semiquantitative estimate of torsional rigidity in the nucleotide backbone.  相似文献   

8.
The dispersion properties of a magnetoactive plasma-filled waveguide in the vicinity of the upper hybrid frequency ω1 = (ω p 2 + ω B 2 )1/21 ? δ < ω < ω1) are studied. It is shown that, in this frequency range, the eigenmodes of the plasma-filled waveguide are represented by the families of EH and cyclotron HE modes, the interaction between which is weak everywhere except for the vicinities of certain points in the (ω, k z ) plane. The equations describing the behavior of the dispersion curves in these vicinities are derived. It is shown that, as a result of the interaction, the high-frequency branches of EH modes acquire the cutoff frequencies corresponding to the high-order propagating HE modes. It is established that the anisotropic HE+l mode can also interact with cyclotron HE modes. In this case, its dispersion curve enters the lower half-vicinity of the upper hybrid frequency, where it is modified due to the interaction, and then leaves it.  相似文献   

9.
10.
Abstract: The aging-associated changes in hippocampal benzodiazepine (ω) receptor isotypes have been investigated in rats of the Wistar and Fischer 344 strains. Displacement experiments of [3H]flunitrazepam binding by zolpidem demonstrated that in hippocampal membranes from adult (3-month-old) Wistar strain rats, high (type I; ω1)-, intermediate (type IIM; ω2)-, and low (type IIL; ω5)- affinity sites for this imidazopyridine account for 27.1 ± 7.5, 44.2 ± 7.5, and 28.8 ± 5.1%, respectively. In hippocampal membranes from aged (24-month-old) rats of the same strain, the relative abundance of these sites was 42.8 ± 9.3, 26.3 ± 4, and 36.0 ± 5.9%, respectively. In contrast, no significant difference was observed in the whole benzodiazepine (ω) binding site density between adult and aged rats. The increase in type I (ω1) binding site density in the hippocampus of aged rats was also demonstrated in saturation experiments with [3H]zolpidem. This aging-induced increase in [3H]zolpidem binding was also observed in hippocampal membranes from Fischer 344 rats. Moreover, in both rat strains, GABA induced a greater enhancement of [3H]zolpidem (5 nM) binding to type I (ω1) sites (GABA shift) in aged than in adult hippocampal membranes. Quantitative autoradiographic analysis of [3H]zolpidem binding to coronal brain sections from adult and aged Fischer 344 rats demonstrated that the aging-associated increases in the density of type I (ω1) binding sites were restricted to the hippocampus. Moreover, increases in binding density were larger in the dentate gyrus and in the CA2 field than in the CA1 and CA3 fields.  相似文献   

11.
A uniform notation and convention is suggested to describe the torsional angles in nucleic acids and their derivatives. The torsional angle χ, relating the stereochemistry of the base with respect to the sugar, shows more variation for the β-purine glycosides than for the β-pyrimidine glycosides. This variation is attributed to the fact that the β-purine derivatives may form intramolecular O(5′)-H…N(3) hydrogen bonding. The χ values for the α-purine and α-pyrimidine glycosides show preference for the –syn-clinal (or anti) conformation. The mode of puckering of the sugar also influences the χ value. The various possible conformations for the furanose ring are described by the torsional angles τ0 τ1, τ2, τ3, τ4, about the five ring bonds. From an analysis of the torsional angles (ω, ?, ψ, ψ′, ?′, ω′) about the sugar phosphate bonds in the x-ray structures of the known nucleosides, nucleotides, phosphodiesters, nucleic acids, and related compounds, and from a consideration of molecular models, it is found that the possible conformations for the backbone of helical nucleic acids is strikingly limited. Most importantly, the preferred conformation of the nucleotide unit in poly nucleotides and nucleic acids turns out to be the same as that found for the nucleotide in the crystal structure. It is observed that base “stacking” is a consequence of the restricted backbone conformation. The torsional angles are illustrated in the form of conformational “wheels”. Interrelation between the torsion angles about successive pairs of sugar-phosphate bonds are presented in the form of conformational maps: ω,?; ?,ψ; ψ.ψ′; ψ′,?′; ?′,ω′; ω′,ω. The ω′,ω map shows the perferred conformations about the inter-nucleotide bonds of right- and left-handed helices and the possible conformations of phosphodiesters. The preferred conformation of the pyrophosphate and triphosphate is that in which the phosphate oxygens display a staggered arrangement when viewed along the P–P axis. A plausible structure and conformation for the ATPM2? backbound complex is presented. This structure differs from that proposed by SzentGyorgi in that the metal (only transition metals are considered here) is not bound to the NH2 nitrogen of adenine, but rather is simultaneously bound to N(7) of the ring and three phosphates (α, β, γ), or N(7) of the ring and two phosphates (β, γ). The remaining metal coordination may be satisfied by solvent–metal or enzyme–metal bonds.  相似文献   

12.
Chemical mutation of a strain producing hydroxy-fatty acid from oleic acid (OA) using NTG (N-methyl-N’-nitro-N-nitrosoguanidine) resulted in a high percentage of improved mutants. A positive screening procedure yielded several high producers, specifically the strain Bacillus pumilus M-F641 (BP M-F641). This strain produced predominantly ω?1-hydroxy fatty acid and could utilize higher concentrations of OA than the parent strain. In shake flask culture, the best ω?1-hydroxy fatty acid concentration and yield (the ratio of ω?1-hydroxy fatty acid accumulation to OA consumption) reached 570 mg L?1 and 11.5%, respectively. Repeated tests showed that the mutant BP M-F641 was genetically stable.  相似文献   

13.
Xie  Fuquan  Pei  Shengxiang  Huang  Xiaoyun  Wang  Lina  Kou  Jinyan  Zhang  Gaiyun 《Antonie van Leeuwenhoek》2021,114(12):2133-2145

A novel Gram-staining positive, aerobic, rod-shaped, non-motile and yellow-pigmented actinobacterium, designated strain WY83T, was isolated from a marine sediment of Indian Ocean. Strain WY83T grew optimally at 30–35 °C, pH 7–8 and with 0–3% (w/v) NaCl. The predominant menaquinones were MK-10, MK-11 and MK-12, and the major fatty acids were C19:1 ω9c/C19:1 ω11c, anteiso-C15:0, C17:0 3OH, and iso-C16:0. The polar lipids consisted of diphosphatidylglycerol, phosphatidylglycerol and one unidentified glycolipid. The cell-wall peptidoglycan contained lysine as a diamino acid. The DNA G?+?C content was 72.3 mol%. Phylogenetic analysis based on 16S rRNA gene sequences and ninety-two bacterial core genes indicated that strain WY83T formed an evolutionary lineage with Chryseoglobus frigidaquae JCM 14730T, Chryseoglobus indicus CTD02-10-2T, Yonghaparkia alkaliphila JCM 15138T, Microcella alkaliphila DSM 18851T and Microcella putealis DSM 19627T within the radiation enclosing members of the family Microbacteriaceae. All pairwise percentage of conserved proteins between strain WY83T and the closely related phylogenetic neighbors were greater than 65%. The average nucleotide identity and in silico DNA–DNA hybridization values were both below the thresholds used for the delineation of a new species. On the basis of the evidence presented, strains WY83T, Y. alkaliphila JCM 15138T, C. frigidaquae JCM 14730T, M. alkaliphila DSM 18851T and M. putealis DSM 19627T should belong to different species of the same genus. Strain WY83T represents a novel species of the genus Microcella, for which the name Microcella flavibacter sp. nov. is proposed. The type strain is WY83T (=?KCTC 39637T?=?MCCC 1A07099T). Furthermore, Chryseoglobus frigidaquae, Chryseoglobus indicus, and Yonghaparkia alkaliphila were reclassified as Microcella frigidaquae comb. nov., Microcella indica nom. nov., and Microcella alkalica nom. nov., respectively.

  相似文献   

14.
A novel Gram-negative, orange-pigmented bacterial strain JLT2008T was isolated from the surface seawater of the Western Pacific and subjected to a polyphasic taxonomic study. Phylogenetic analysis based on 16S rRNA gene sequences revealed that strain JLT2008T belonged to the genus Erythrobacter, sharing the highest similarity (96.6 %) with Erythrobacter gangjinensis K7-2T and the lowest similarity (94.9 %) with Erythrobacter litoralis DSM 8509T. Strain JLT2008T did not contain bacteriochlorophyll a, and the predominant respiratory lipoquinone was ubiquinone-10. The major fatty acids were C18:1 ω7c, C16:0, C16:1 ω7c/C16:1 ω6c. The prominent polar lipids were sphingoglycolipid, phosphatidylethanolamine, and phosphatidylglycerol. The genomic G + C content was 60.1 mol %. Based on the polyphasic taxonomic data, a novel species within the genus Erythrobacter, and with the name Erythrobacter westpacificensis sp. nov., is proposed. The type strain is JLT2008T (=CGMCC 1.10993T = JCM 18014T).  相似文献   

15.
Two novel aerobic p-n-nonylphenol-degrading bacterial strains were isolated from seawater obtained from the coastal region of Ogasawara Islands, Japan. The 16S rRNA gene sequence analysis indicated that the strains are affiliated with the order Alteromonadales within the class Gammaproteobacteria. One isolate, strain KU41G2, is most closely related to Maricurvus nonylphenolicus (99.2 % similarity), and is tentatively identified as M. nonylphenolicus. The other isolate, strain KU41GT, is also most closely related to M. nonylphenolicus; however, the 16S rRNA gene sequence similarity was only 94.7 %. Cells of strain KU41GT are Gram-negative rods with a single polar flagellum. The predominant respiratory lipoquinone was ubiquinone-8, and the major cellular fatty acids were C17:1 ω8c (24.2 %); C15:0 iso 2-OH; and/or C16:1 ω7c (16.3 %), C15:0 (10.3 %), C11:0 3-OH (9.5 %), C9:0 3-OH (6.7 %), C10:0 3-OH (6.4 %), and C18:1 ω7c (5.5 %). The DNA G+C content was 53.3 mol%. On the basis of physiological, chemotaxonomic, and phylogenetic data, strain KU41GT is suggested to represent a novel species of a new genus, for which we propose the name Pseudomaricurvus alkylphenolicus gen. nov., sp. nov. The type strain of P. alkylphenolicus is KU41GT (=JCM 19135T = KCTC 32386T).  相似文献   

16.
Turnover of the ubiquinol oxidizing site of the UQH2:cyt c2 oxidoreductase (b/c 1 complex) ofRps. sphaeroides can be assayed by measuring the rate of reduction of cytb 561 in the presence of antimycin (AA). Oxidation of ubiquinol is a second-order process, with a value ofk 2 of about 3 × 105 M–1. The reaction shows saturation at high quinol concentrations, with an apparentK m of about 6–8 mM (with respect to the concentration of quinol in the membrane). When the quinone pool is oxidized before illumination, reduction of the complex shows a substantial lag (about 1 ms) after a flash, indicating that the quinol produced as a result of the photochemical reactions is not immediately available to the complex. We have suggested that the lag may be due to several factors, including the leaving time of the quinol from the reaction center, the diffusion time to the complex, and the time for the head group to cross the membrane. We have suggested aminimal value for the diffusion coefficient of ubiquinone in the membrane (assuming that the lag is due entirely to diffusion) of about 10–9 cm–2 sec–1. The lag is reduced to about 100 µsec when the pool is significantly reduced, showing that quinol from the pool is more rapidly available to the complex than that from the reaction center. With the pool oxidized, similar kinetics are seen when the reduction of cytb 561 occurs through the AA-sensitive site (with reactions at the quinol oxidizing site blocked by myxothiazol). These results show that there is no preferential reaction pathway for transfer of reducing equivalents from reaction center tob/c 1 complex. Oxidation of cytb 561 through the AA-sensitive site can be assayed from the slow phase of the carotenoid electrochromic change, and by comparison with the kinetics of cytb 561. As long as the quinone pool is significantly oxidized, the reaction is not rate-determining for the electrogenic process. On reduction of the pool below 1 quinone per complex, a slowing of the electrogenic process occurs, which could reflect a dependence on the concentration of quinone. If the process is second-order, the rate constant must be about 2–5 times greater than that for quinol oxidation, since the effect on rate is relatively small compared with the effect seen at the quinol oxidizing site when the quinol concentration is changed over theE h range where the first few quinols are produced on reductive titration. When the quinone pool is extracted (experiments in collaboration with G. Venturoli and B. A. Melandri), the slowing of the electrochromic change on reduction of the pool is not enhanced; we assume that this is due to the fact that a minimum of one quinone per active complex is produced by turnover of the quinol oxidizing site. Two lines of research lead us to revise our previous estimate for the minimal value of the quinone diffusion coefficient. These relate to the relation between the diffusion coefficient and the rate constants for processes involving the quinones: (a) The estimated rate constant for reaction of quinone at the AA-site approaches the calculated diffusion limited rate constant, implying an improbably efficient reaction. (b) From a preliminary set of experiments, the activation energy determined by measuring the variation of the rate constant for quinol oxidation with temperature, is about 8 kcal mol–1. Although we do not know the contribution of entropic terms to the pre-exponential factor, the result is consistent with a considerably larger value for the diffusion coefficient than that previously suggested.  相似文献   

17.
Moxonidine, an imidazoline deriviatives, suppress the vasopressor sympathetic outflow to produce hypotension. This effect has been known to be mediated in part by suppressing sympathetic outflow via acting imidazoline I1 receptors (IR1) at postganglionic sympathetic neurons. But, the cellular mechanism of IR1-induced inhibition of noradrenaline (NA) release is still unknown. We therefore, investigated the effect of IR1 activation on voltage-dependent Ca2+ channels which is known to play an pivotal role in regulating NA in rat superior cervical ganglion (SCG) neurons, using the conventional whole-cell patch-clamp method. In the presence of rauwolscine (3 μΜ), which blocks α2-adrenoceptor (Rα2), moxonidine inhibited voltage-dependent Ca2+ current (ICa) by about 30%. This moxonidine-induced inhibition was almost completely prevented by efaroxan (10 μΜ) which blocks IR1 as well as Rα2. In addition, ω-conotoxin (CgTx) GVIA (1 μΜ) occluded moxonidine-induced inhibition of ICa, but, moxonidine-induced ICa inhibition was not affected by pertussis toxin (PTX) nor shows any characteristics of voltage-dependent inhibition. These data suggest that moxonidine inhibit voltage-dependent N-type Ca2+ current (ICa–N) via activating IR1. Finally, moxonidine significantly decreased the frequency of AP firing in a partially reversible manner. This inhibition of AP firing was almost completely occluded in the presence of ω-CgTx. Taken together, our results suggest that activation of IR1 in SCG neurons reduced ICa–N in a PTX-and voltage-insensitive pathway, and this inhibition attenuated repetitive AP firing in SCG neurons.  相似文献   

18.
1. Analyses of zooplankton fatty acid (FA) composition in laboratory experiments and samples collected from lakes in New Zealand spanning a wide gradient of productivity were used to assess the extent to which FAs might infer their diet. We used the cladocerans, Daphnia and Ceriodaphnia, and the calanoid copepod, Boeckella, as test organisms, and monocultures of cryptophytes, chlorophytes and cyanobacteria as food. Based on reproductive success, cryptophytes were the highest food quality, chlorophytes were intermediate and cyanobacteria the poorest. 2. Several FA groups were highly correlated between zooplankton and their diets. They were monounsaturated fatty acids (MUFAs), and ω3 and ω6 polyunsaturated fatty acids (PUFAs) for cladocerans, and saturated fatty acids (SAFAs) and ω3 PUFAs for copepods. Several FAs varied significantly less in the zooplankton than in their monoculture diets, e.g. MUFAs in Daphnia, and ω3 and ω6 PUFAs in Ceriodaphnia, despite clear dietary dependency for these FAs. 3. Zooplankton collected from lakes in New Zealand had more eicosapentaenoic acid (EPA) (Daphnia), more highly unsaturated ω3 and ω6 FAs (C20, C22; Daphnia, Ceriodaphnia, Boeckella) and less ω3 C18 PUFAs (Daphnia, Ceriodaphnia, Boeckella) and ω6 C18 PUFAs (Daphnia, Ceriodaphnia) than measured in the same species reared on phytoplankton in the laboratory. 4. Analyses of FA composition of seston and freshwater zooplankton globally showed that, in general, zooplankton had a significantly higher proportion of arachidonic acid and EPA than seston, and copepods also had a higher percentage of docosahexaenoic acid than seston. 5. These results suggest that zooplankton selectively incorporate the most physiologically important FAs. This could be a consequence of preferential assimilation, selective feeding on more nutritious cells or locating and feeding within higher food quality food patches.  相似文献   

19.
Abstract: LAN-1 is a human neuroblastoma cell line that, in the undifferentiated state, does not respond to membrane depolarization with an elevation of [Ca2+]i, monitored by fura-2 single-cell microfluorimetry. The exposure of LAN-1 cells to the differentiating agent retinoic acid induced the appearance of [Ca2+]i elevation elicited by 55 mM K+. Maitotoxin, a putative activator of voltage-sensitive Ca2+ channels, did not evoke an elevation of [Ca2+]i in undifferentiated LAN-1 cells, but produced a marked and sustained increase in [Ca2+]i when superfused in retinoic acid-treated cells. Both high K+- and maitotoxin-induced [Ca2+]i elevation in retinoic acid-differentiated LAN-1 cells was reversed by the lanthanide Gd3+, an inorganic Ca2+-entry blocker, and by the snail toxin ω-conotoxin GVIA, which interacts with the N sub-type of voltage-sensitive Ca2+ channels. In contrast, both Bay K 8644 and nimodipine, dihydropyridines that selectively activate or block, respectively, the L-channel sub-type, were completely ineffective. The tumor promoter phorbol 12-myristate 13-acetate (100 nM), a protein kinase C activator, inhibited the elevation of [Ca2+]i due to Ca2+ influx elicited by membrane depolarization. K+-induced [Ca2+]i elevation appeared 24 h after the addition of retinoic acid and reached the highest magnitude after 72 h. Furthermore, 8 days after the removal of the differentiating agent from the culture medium, the high K+-induced increase of [Ca2+]i was still present. In conclusion, the results of the present study demonstrated that retinoic acid-induced differentiation of LAN-1 cells, which lack a high K+-evoked [Ca2+]i increase in the undifferentiated state, induces the functional expression of an ω-conotoxin GVIA-sensitive, dihydropyridine-insensitive N-type voltage-sensitive Ca2+ channel that can be activated by maitotoxin and negatively modulated by protein kinase C.  相似文献   

20.
A novel endophytic actinomycete, designated strain NEAU-J3T, was isolated from soybean root (Glycine max (L.) Merr) and characterized using a polyphasic approach. Phylogenetic analysis based on 16S rRNA gene sequences suggested that strain NEAU-J3T fell within the family Micromonosporaceae. The strain was observed to form an extensively branched substrate mycelium, which carried non-motile oval spores with a smooth surface. The cell walls of strain NEAU-J3T were determined to contain meso-diaminopimelic acid and galactose, ribose and glucose were detected as whole-cell sugars. The major menaquinones were determined to be MK-9(H4) and MK-9(H6). The phospholipids detected were phosphatidylcholine and phosphatidylethanolamine. The major cellular fatty acids were determined to be C16:0, C18:1 ω9c, C18:0, C17:0, C17:1 ω7c, anteiso-C17:0, C16:1 ω7c and C15:0. The DNA G + C content was 62.5 mol%. On the basis of the morphological and chemotaxonomic characteristics, phylogenetic analysis and characteristic patterns of 16S rRNA gene signature nucleotides, strain NEAU-J3T is considered to represent a novel species of a new genus within the family Micromonosporaceae, for which the name Wangella harbinensis gen. nov., sp. nov. is proposed. The type strain of Wangella harbinensis is strain NEAU-J3T (=CGMCC 4.7039T = DSM 45747T).  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