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 328 毫秒
1.
PvuRts1I is a modification-dependent restriction endonuclease that recognizes 5-hydroxymethylcytosine (5hmC) as well as 5-glucosylhydroxymethylcytosine (5ghmC) in double-stranded DNA. Using PvuRts1I as the founding member, we define a family of homologous proteins with similar DNA modification-dependent recognition properties. At the sequence level, these proteins share a few uniquely conserved features. We show that these enzymes introduce a double-stranded cleavage at the 3'-side away from the recognized modified cytosine. The distances between the cleavage sites and the modified cytosine are fixed within a narrow range, with the majority being 11-13 nt away in the top strand and 9-10 nt away in the bottom strand. The recognition sites of these enzymes generally require two cytosines on opposite strand around the cleavage sites, i.e. 5'-CN(11-13)↓N(9-10)G-3'/3'-GN(9-10)↓N(11-13)C-5', with at least one cytosine being modified for efficient cleavage. As one potential application for these enzymes is to provide useful tools for selectively mapping 5hmC sites, we have compared the relative selectivity of a few PvuRts1I family members towards different forms of modified cytosines. Our results show that the inherently different relative selectivity towards modified cytosines can have practical implications for their application. By using AbaSDFI, a PvuRts1I homolog with the highest relative selectivity towards 5ghmC, to analyze rat brain DNA, we show it is feasible to map genomic 5hmC sites close to base resolution. Our study offers unique tools for determining more accurate hydroxymethylomes in mammalian cells.  相似文献   

2.
FACE水稻茎蘖动态模型   总被引:4,自引:0,他引:4  
借助农田开放式空气CO2浓度增高(FACE)技术平台,以武香粳14为供试水稻品种,设置不同施N量处理,研究大气CO2浓度为570 μmol·mol-1(比对照高200 μmol·mol-1)的FACE处理对水稻茎蘖动态的影响,并建立了相应的模拟模型:Tt=A1(1+ea1-b1t)-A2(1+ea2-b2t)+C×[B1(1+ea3-b3t)-B2(1+ea4-b4t)]+D.模型以时间为驱动因子,描述了水稻茎蘖数随移栽天数的动态变化过程,对常规及CO2浓度增加条件下水稻茎蘖的变化均有很好的拟合性.通过不同年份试验数据对模型的检验,预测根均方差(RMSE)最大为44.27个·m-2,最小为13.96个·m-2,且相关系数均达到了极显著水平.表明模型的预测程度较高,具有很好的适用性.  相似文献   

3.
Lee I  Deng W  Yang L  Wang C  Bai C 《Biophysical chemistry》1997,67(1-3):159-165
The conformational transitions (helix-coil transitions) of three hairpin triple helices, models 5'-(A-G)(3) + 5'-(T-C)(3)-T(4)-((br)C-T)(3) [CY], 5'-(A-G)(3) + 5'-(T-(br)C)(3)-T(4)-(C-T)(3) [YC] and 5'-(A-G)(3) + 5'-(T-(br)C)(3)-T(4)-((br)C-T)(3) [YY], are characterized in this work by UV spectroscopy. Melting of these triplexes is biphasic, and the profiles are used to obtain the thermodynamic parameters. The thermodynamic properties of the hairpin triplex are T(m) = 19.45 degrees C and DeltaH(vH) = 293.12 kJ mol(-1) for CY, T(m) = 22.85 degrees C and DeltaH(vH) = 256.63 kJ mol(-1) for YC and T(m) = 28.47 degrees C and DeltaH(vH) = 234.68 kJ mol(-1) for YY at pH 4.4. Those of the duplex are T(m) = 30.50 degrees C and DeltaH(vH) = 427.09 kJ mol(-1) for CY, T(m) = 32.96 degrees C and DeltaH(vH) = 374.47 kJ mol(-1) for YC and T(m) = 33.24 degrees C and DeltaH(vH) = 329.67 kJ mol(-1) for YY at pH 4.4. The distinct transitions of triplex to duplex and duplex to single strands are analyzed using the nearest-neighbor Ising model. Electrostatic effects on each conformation are also analyzed.  相似文献   

4.
Simmen MW 《Genomics》2008,92(1):33-40
In mammalian genomes CpGs occur at one-fifth their expected frequency. This is accepted as resulting from cytosine methylation and deamination of 5-methylcytosine leading to TpG and CpA dinucleotides. The corollary that a CpG deficit should correlate with TpG excess has not hitherto been systematically tested at a genomic level. I analyzed genome sequences (human, chimpanzee, mouse, pufferfish, zebrafish, sea squirt, fruitfly, mosquito, and nematode) to do this and generally to assess the hypothesis that CpG deficit, TpG excess, and other data are accountable in terms of 5-methylcytosine mutation. In all methylated genomes local CpG deficit decreases with higher G + C content. Local TpG surplus, while positively associated with G + C level in mammalian genomes but negatively associated with G + C in nonmammalian methylated genomes, is always explicable in terms of the CpG trend under the methylation model. Covariance of dinucleotide abundances with G + C demonstrates that correlation analyses should control for G + C. Doing this reveals a strong negative correlation between local CpG and TpG abundances in methylated genomes, in accord with the methylation hypothesis. CpG deficit also correlates with CpT excess in mammals, which may reflect enhanced cytosine mutation in the context 5'-YCG-3'. Analyses with repeat-masked sequences show that the results are not attributable to repetitive elements.  相似文献   

5.
6.
Liu Q  Wen CK 《Plant physiology》2012,158(3):1193-1207
The ethylene response is negatively regulated by a family of five ethylene receptor genes in Arabidopsis (Arabidopsis thaliana). The five members of the ethylene receptor family can physically interact and form complexes, which implies that cooperativity for signaling may exist among the receptors. The ethylene receptor gene mutations etr1-1((C65Y))(for ethylene response1-1), ers1-1((I62P)) (for ethylene response sensor1-1), and ers1(C65Y) are dominant, and each confers ethylene insensitivity. In this study, the repression of the ethylene response by these dominant mutant receptor genes was examined in receptor-defective mutants to investigate the functional significance of receptor cooperativity in ethylene signaling. We showed that etr1-1((C65Y)), but not ers1-1((I62P)), substantially repressed various ethylene responses independent of other receptor genes. In contrast, wild-type receptor genes differentially supported the repression of ethylene responses by ers1-1((I62P)); ETR1 and ETHYLENE INSENSITIVE4 (EIN4) supported ers1-1((I62P)) functions to a greater extent than did ERS2, ETR2, and ERS1. The lack of both ETR1 and EIN4 almost abolished the repression of ethylene responses by ers1(C65Y), which implied that ETR1 and EIN4 have synergistic effects on ers1(C65Y) functions. Our data indicated that a dominant ethylene-insensitive receptor differentially repressed ethylene responses when coupled with a wild-type ethylene receptor, which supported the hypothesis that the formation of a variety of receptor complexes may facilitate differential receptor signal output, by which ethylene responses can be repressed to different extents. We hypothesize that plants can respond to a broad ethylene concentration range and exhibit tissue-specific ethylene responsiveness with differential cooperation of the multiple ethylene receptors.  相似文献   

7.
The effect of temperature (between 5.0 and 45.0 degrees C) on the catalytic activity of nitric oxide synthases types I, II, and III (NOS-I, NOS-II, and NOS-III, respectively) has been investigated, at pH 7.5. The value of V(max) for NOS-I activity increases from 1.8 x 10(1) pmol min(-1) mg(-1), at 5.0 degrees C, to 1.8 x 10(2) pmol min(-1) mg(-1), at 45.0 degrees C; on the other hand, the value of K(m) (=4.0 x 10(-6) M) is temperature independent. Again, the value of V(max) for NOS-II activity increases from 8.0 pmol min(-1) mg(-1), at 7.0 degrees C, to 5.4 x 10(1) pmol min(-1) mg(-1), at 40.0 degrees C, the value of K(m) (=1.8 x 10(-5) M) being unaffected by temperature. Temperature exerts the same effect on NOS-I and NOS-II activity, as shown by the same values of DeltaH(V(max)) (=4.2 x 10(1) kJ mol(-1)), DeltaH(K(m)) (=0 kJ mol(-1)), and DeltaH((V(max))(/K(m))()) (=4.2 x 10(1) kJ mol(-1)). On the contrary, the value of K(m) for NOS-III activity decreases from 3.8 x 10(-5) M, at 10.0 degrees C, to 1.6 x 10(-5) M, at 40.0 degrees C, the value of V(max) (=6.8 x 10(1) pmol min(-1) mg(-1)) being temperature independent. Present results indicate that temperature influences directly NOS-I and NOS-II activity independently of the substrate concentration, the values of K(m) being temperature independent. However, when l-arginine level is higher than 2 x 10(-4) M, as observed under in vivo conditions, NOS-III activity is essentially unaffected by temperature, the substrate concentration exceeding the value of K(m). As a whole, although further studies in vivo are needed, these observations seem to have potential physiopathologic implications.  相似文献   

8.
The biosyntheses of aroma active gamma- and delta-lactones have been previously characterized in yeasts and plants by incubation of labeled fatty acid derivatives. The lactones were considered as end products. Liquid cultures of the lactone-producing yeast Sporobolomyces odorus were used to investigate catabolic pathways of the lactones by incubation of ethyl (+/-)-5-hydroxy(1-(13)C1)decanoate ((13C)-1b) and methyl (+/-)-4-hydroxy(1-(13)C1)decanoate ((13C)-7a). Aliquots of the culture broth were analyzed with GC/MS after CH2N2 derivatization. S. odorus degraded (13C)-1b to 5-oxo(1-(13)C1)decanoic acid ((13C)-2c) and, subsequently, to pentyl (1-(13)C1)pentanedioate ((13C)-3c) and 3-[(1-(13)C1)carboxypropyl] hexanoate ((13C)-4c) by a Baeyer-Villiger-type oxidation (BVO). In addition, the oxidation of (13C)-7a to 4-oxo(1-(13)C1)decanoic acid ((13C)-8c) and a BVO of (13C)-8c to hexyl (1-(13)C1)butanedioate ((13C)-9c) is reported. So far, BVO has been observed in bacteria and some fungi; the data presented indicate a BVO catalyzed by the yeast S. odorus in the course of endogenous lactone metabolism.  相似文献   

9.
The epigenetic DNA modifications 5-methylcytosine (5mC) and 5-hydroxymethylcytosine (5hmC) in eukaryotes are recognized either in the context of double-stranded DNA (e.g., by the methyl-CpG binding domain of MeCP2), or in the flipped-out state (e.g., by the SRA domain of UHRF1). The SRA-like domains and the base-flipping mechanism for 5(h)mC recognition are also shared by the recently discovered prokaryotic modification-dependent endonucleases of the MspJI and PvuRts1I families. Since the mechanism of modified cytosine recognition by many potential eukaryotic and prokaryotic 5(h)mC “readers” is still unknown, a fast solution based method for the detection of extrahelical 5(h)mC would be very useful. In the present study we tested base-flipping by MspJI- and PvuRts1I-like restriction enzymes using several solution-based methods, including fluorescence measurements of the cytosine analog pyrrolocytosine and chemical modification of extrahelical pyrimidines with chloroacetaldehyde and KMnO4. We find that only KMnO4 proved an efficient probe for the positive display of flipped out pyrimidines, albeit the method required either non-physiological pH (4.3) or a substitution of the target cytosine with thymine. Our results imply that DNA recognition mechanism of 5(h)mC binding proteins should be tested using a combination of all available methods, as the lack of a positive signal in some assays does not exclude the base flipping mechanism.  相似文献   

10.
A new protocol for superovulating cattle which allows for control of the timing of ovulation after superstimulation with FSH was developed. The preovulatory LH surge was blocked with the GnRH agonist deslorelin, and ovulation was induced by injection of LH. In Experiment 1, heifers (3-yr-old) were assigned to a control group (Group 1A, n = 4) or a group with deslorelin implants (Group 1B, n = 5). On Day -7, heifers in Group 1A received a progestagen CIDR-B((R))device, while heifers in Group 1B received a CIDR-B((R))device + deslorelin implants. Both groups were superstimulated with twice daily injections of FSH (Folltropin((R))-V): Day 0, 40 mg (80 mg total dose on Day 0); Day 1, 30 mg; Day 2, 20 mg; Day 3, 10 mg. On Day 2, heifers were given PGF (a.m.) and CIDR-B((R)) devices were removed (p.m.). Three heifers in Group 1A had a LH surge and ovulated, whereas neither of these events occurred in Group 1B (with deslorelin implants) heifers. In Experiment 2, heifers (3-yr-old) were assigned to 1 of 4 equal groups (n = 6). On Day -7, heifers in Group 2A received a norgestomet implant, while heifers in Groups 2B, 2C and 2D received norgestomet + deslorelin implants. Heifers were superstimulated with FSH starting on Day 0 as in Experiment 1. On Day 2, heifers were given PGF (a.m.) and norgestomet implants were removed (p.m.). Heifers in Groups 2B to 2D were given 25 mg LH (Lutropin((R))): Group 2B, Day 4 (a.m.); Group 2C, Day 4 (p.m.); Group 2D, Day 5 (a.m.). Heifers in Group 2A were inseminated at estrus and 12 and 24 h later, while heifers in Groups 2B to 2D were inseminated at the time of respective LH injection and 12 and 24 h later. Injection of LH induced ovulation in heifers in Groups 2B to 2D. Heifers in Group 2C had similar total ova and embryos (15.2 +/- 1.4) as heifers in Group 2A (11.0 +/- 2.8) but greater (P < 0.05) numbers than heifers in Group 2B (7.0 +/- 2.3) and Group 2D (6.3 +/- 2.0). The number of transferable embryos was similar for heifers in Group 2A (5.8 +/- 1.8) and Group 2C (7.3 +/- 2.1) but lower (P < 0.05) for heifers in Group 2B (1.2 +/- 0.8) and Group 2D (1.3 +/- 1.0). The new GnRH agonist-LH protocol does not require observation of estrus, and induces ovulation in superstimulated heifers that would not have an endogenous LH surge.  相似文献   

11.
The mechanisms by which DNA-incorporated radionuclides impart lethal damage to mammalian cells were investigated by examining the capacity of dimethyl sulfoxide (DMSO) to protect against lethal damage to Chinese hamster V79 cells caused by unbound tritium ((3)H(2)O), DNA-incorporated (125)I- and (131)I-iododeoxyuridine ((125)IdU, (131)IdU), and cytoplasmically localized (210)Po citrate. The radionuclides (3)H and (131)I emit low- and medium-energy beta particles, respectively, (125)I is a prolific Auger electron emitter, and (210)Po emits 5.3 MeV alpha particles. Cells were radiolabeled and maintained at 10.5 degrees C for 72 h in the presence of different concentrations of DMSO (5-12.5% v/v), and the surviving fraction compared to that of unlabeled controls was determined. DMSO afforded no protection against the lethal effects of the high-LET alpha particles emitted by (210)Po. Protection against lethal damage caused by unbound (3)H, (131)IdU and (125)IdU depended on the concentration of DMSO in the culture medium. Ten percent DMSO provided maximum protection in all cases. The dose modification factors obtained at 10% DMSO for (3)H(2)O, (131)IdU, (125)IdU and (210)Po citrate were 2.9 +/- 0.01, 2.3 +/- 0.5, 2.6 +/- 0.2 and 0.95 +/- 0.07, respectively. These results indicate that the toxicity of Auger electron and beta-particle emitters incorporated into the DNA of mammalian cells is largely radical-mediated and is therefore indirect in nature. This is also the case for the low-energy beta particles emitted by (3)H(2)O. In contrast, alpha particles impart lethal damage largely by direct effects. Finally, calculations of cellular absorbed doses indicate that beta-particle emitters are substantially more toxic when incorporated into the DNA of mammalian cells than when they are localized extracellularly.  相似文献   

12.
N Zhang  J E Womack 《Genomics》1992,14(1):126-130
In an effort to generate a more complete bovine syntenic map of Type I comparative anchor loci, seven homologs to genes found on HSA5 were mapped using a panel of bovine x rodent hybrid somatic cells. Five HSA5 genes, CSF2, RPS14, PDGFRB, FGFA, and CSF1R, were assigned to bovine syntenic group U22 (chromosome 7), while two others, C9 and HGMCR, mapped to U10 and U5, respectively. Previous studies had assigned the HSA5 marker SPARC to bovine syntenic group U22. The mapping of genes spanning the length of HSA5 in cattle and also in mouse permits syntenic comparisons between prototypic genomes of three mammalian orders, providing insight into the evolutionary history of this region of the ancestral mammalian genome.  相似文献   

13.
Human cellular DNA contains two distinguishable families of retroviral related sequences. One family shares extensive nucleotide sequence homology with infectious mammalian type C retroviral genomes (T. I. Bonner, C. O'Connell, and M. Cohen, Proc. Natl. Acad. Sci. USA 79:4709-4713, 1982; M. A. Martin, T. Bryan, S. Rasheed, and A. S. Khan, Proc. Natl. Acad. Sci. USA 78:4892-4896, 1981). The other family contains major regions of homology with the pol genes of infectious type A and B and avian type C and D retroviral genomes (R. Callahan, W. Drohan, S. Tronick, and J. Schlom, Proc. Natl. Acad. Sci. USA 79:5503-5507, 1982; I. M. Chiu, R. Callahan, S. R. Tronick, J. Schlom, and S. A. Aaronson, Science 223:364-370, 1984). Analysis of the human recombinant clone HLM-2 has shown that the pol gene in the latter family is located within an endogenous proviral genome (R. Callahan, I. M. Chiu, J. F. H. Wong, S. R. Tronick, B. A. Roe, S. A. Aaronson, and J. Schlom, Science 228:1208-1211, 1985). We show that the proviral genome in HLM-2 and the related recombinant clone HLM-25 are located, respectively, on human chromosomes 1 and 5. Other related proviral genomes are located on chromosomes 7, 8, 11, 14, and 17.  相似文献   

14.
Glyoxalase I from yeast (Saccharomyces cerevisiae) purified by affinity chromatography on S-hexylglutathione-Sepharose 6B was characterized and compared with the enzyme from rat liver, pig erythrocytes and human erythrocytes. The molecular weight of glyoxalase I from yeast was, like the enzyme from Rhodospirillum rubrum and Escherichia coli, significantly less (approx. 32000) than that of the enzyme from mammals (approx. 46000). The yeast enzyme is a monomer, whereas the mammalian enzymes are composed of two very similar or identical subunits. The enzymes contain 1Zn atom per subunit. The isoelectric points (at 4 degrees C) for the yeast and mammalian enzymes are at pH7.0 and 4.8 respectively; tryptic-peptide ;maps' display corresponding dissimilarities in structure. These and some additional data indicate that the microbial and the mammalian enzymes may have separate evolutionary origins. The similarities demonstrated in mechanistic and kinetic properties, on the other hand, indicate convergent evolution. The k(cat.) and K(m) values for the yeast enzyme were both higher than those for the enzyme from the mammalian sources with the hemimercaptal adduct of methylglyoxal or phenylglyoxal as the varied substrate and free glutathione at a constant and physiological concentration (2mm). Glyoxalase I from all sources investigated had a k(cat.)/K(m) value near 10(7)s(-1).m(-1), which is close to the theoretical diffusion-controlled rate of enzyme-substrate association. The initial-velocity data show non-Michaelian rate saturation and apparent non-linear inhibition by free glutathione for both yeast and mammalian enzyme. This rate behaviour may have physiological importance, since it counteracts the effects of fluctuations in total glutathione concentrations on the glyoxalase I-dependent metabolism of 2-oxoaldehydes.  相似文献   

15.
X-ray analyses have shown that the glucopyranose rings of GlcNAc-Asn [4-N-(2-acetamido-2-deoxy-beta-d-glucopyranosyl)-l-asparagine] and Glc-Asn [4-N-(beta-d-glucopyranosyl)-l-asparagine] both have the C-1 chair conformation and also that the glucose-asparagine linkage of each molecule is present in the beta-anomeric configuration. The dimensions (the estimated standard deviations of the last digit are in parentheses) of the glycosidic bond in GlcNAc-Asn and Glc-Asn are, respectively, C((1))-N((1)) 0.1441(6)nm, 0.146(2)nm; angle O((5))-C((1))-N((1)) 106.8(3) degrees , 105.7(8) degrees ; angle C((2))-C((1))-N((1)) 111.1(4) degrees , 110.4(9) degrees ; angle C((1))-N((1))-C((9)) 121.4(4) degrees , 120.5(9) degrees . The glycosidic torsion angle C((9))-N((1))-C((1))-C((2)) is 141.0 degrees and 157.6 degrees in GlcNAc-Asn and Glc-Asn respectively. Hydrogen-bonding is extensive in these two crystal structures and does affect one torsion angle in particular. Two very different values of chi(1)(N-C(alpha)-C(beta)-C(gamma)) occur for the asparagine residue of the two different molecules; the values of chi(1), -69.0 degrees in GlcNAc-Asn and 61.9 degrees in Glc-Asn, correspond to two different staggered conformations about the C(alpha)-C(beta) bond as the NH(3) (+) group is adjusted to different hydrogen-bonding patterns. The two trans-peptide groups in GlcNAc-Asn show small distortions in planarity whereas that in Glc-Asn is more non-planar. The mean plane through the atoms of the amide group at C((2)) in GlcNAc-Asn is approximately perpendicular (69 degrees ) to the mean plane through the C((2)), C((3)), C((5)) and O((5)) atoms of the glucose ring and that at C((1)) is less perpendicular (65 degrees ). The mean plane through the atoms of the amide group in Glc-Asn makes an angle of only 55 degrees with the mean plane through these same four atoms of the glucose ring. The N((1))-H bond of the amide at C((1)) is trans to the C((1))-H bond in these two compounds; the N((2))-H bond of the amide at C((2)) is trans to the C((2))-H bond in GlcNAc-Asn. The values of the observed and final calculated structure amplitudes have been deposited as Supplementary Publication SUP 50035 (26 pages) at the British Library (Lending Division), (formerly the National Lending Library for Science and Technology), Boston Spa, Yorks. LS23 7BQ, U.K., from whom copies may be obtained on the terms given in Biochem. J. (1973) 131, 5.  相似文献   

16.
DNA methylation is widespread amongst eukaryotes and prokaryotes to modulate gene expression and confer viral resistance. 5-Methylcytosine (m5C) methylation has been described in genomes of a large fraction of bacterial species as part of restriction-modification systems, each composed of a methyltransferase and cognate restriction enzyme. Methylases are site-specific and target sequences vary across organisms. High-throughput methods, such as bisulfite-sequencing can identify m5C at base resolution but require specialized library preparations and single molecule, real-time (SMRT) sequencing usually misses m5C. Here, we present a new method called RIMS-seq (rapid identification of methylase specificity) to simultaneously sequence bacterial genomes and determine m5C methylase specificities using a simple experimental protocol that closely resembles the DNA-seq protocol for Illumina. Importantly, the resulting sequencing quality is identical to DNA-seq, enabling RIMS-seq to substitute standard sequencing of bacterial genomes. Applied to bacteria and synthetic mixed communities, RIMS-seq reveals new methylase specificities, supporting routine study of m5C methylation while sequencing new genomes.  相似文献   

17.
To gain further insight into the structure/function relationship of cytochrome P450 side-chain cleavage (CYP11A1), this enzyme was investigated in the Cape baboon (Papio ursinus). Four constructs were cloned and characterised in non-steroidogenic mammalian COS-1 cells. Wild type recombinant baboon CYP11A1 cDNA yielded a K(m) value of 1.6 microM for 25-hydroxycholesterol. The single amino acid substitutions, I98Q and I98K resulted in a 1.7- and 2.8-fold increases in K(m) values, respectively. Conversely, the introduction of the mutation, K103A, resulted in a 1.8-fold decrease in K(m). A homology model of CYP11A1, based on the crystal structures of CYP102 and CYP2C5, revealed that residues 98 and 103 lie within the B'-C loop and contribute to the spatial orientation and structural integrity of this domain. Based on these results we propose a topological model of the CYP11A1 active pocket, which is supported by substrate docking analysis and kinetic studies.  相似文献   

18.
1. The identity of the organism previously known as Vibrio O1 (N.C.I.B. 8250) with a species of Moraxella is established. 2. The ability of cells to oxidize propionate is present only in cells with an endogenous respiration and this ability is increased 80-fold when the organism is grown with propionate. 3. Isocitrate lyase activity in extracts from propionate-grown cells is the same as that in extracts from lactate-grown cells, about tenfold greater than that in extracts from succinate-grown cells and slightly greater than half the activity in extracts from acetate-grown cells. 4. With arsenite as an inhibitor conditions were found in which the organism would catalyse the quantitative oxidation of propionate to pyruvate. When propionate was completely utilized pyruvate was metabolized further to 2-oxoglutarate. 5. The oxidation of propionate by cells was incomplete both in a ;closed system' with alkali to trap respiratory carbon dioxide and in an ;open system' with an atmosphere of oxygen+carbon dioxide (95:5). Acetate accumulated. Under these conditions [2-(14)C]- and [3-(14)C]-propionate gave rise to [(14)C]acetate. The rate of conversion of [2-(14)C]propionate into (14)CO(2), although much less than the rate of conversion of [1-(14)C]propionate into (14)CO(2), was slightly greater than the rate of conversion of [3-(14)C]propionate into (14)CO(2). 6. The oxidation of propionate by cells was complete in an ;open system' with an atmosphere of either oxygen or air. Under these conditions very little [1-(14)C]propionate was converted into (14)C-labelled cell material. The conversion of [2-(14)C]- and [3-(14)C]-propionate into (14)C-labelled cell material occurred at an appreciable rate, the rate for the incorporation of [3-(14)C]propionate being slightly more rapid. In the absence of a utilizable nitrogen source part of the [(14)C]propionate was incorporated into some reserve material, which was oxidized when added substrate had been completely utilized. 7. [(14)C]-Pyruvate produced from [(14)C]propionate was chemically degraded. The C((1)) of propionate was found only in C((1)) of pyruvate. At least 86% of C((2)) of pyruvate was derived from C((2)) of propionate and at least 92% of C((3)) of pyruvate from C((3)) of propionate. 8. These results are incompatible with the operation of any of the previously described pathways for propionate metabolism except the direct one, perhaps via an activated acrylate.  相似文献   

19.
The secondary structural features in the 70S RNAs of the Prague strain of avian Rous sarcoma virus, subgroup A (PR-RSV-A), and Moloney murine leukemia virus (M-MuLV) were compared by electron microscopy. The PR-RSV-A genome contained two subunits joined by a linkage structure as in the genomes of M-MuLV and other mammalian retroviruses. In both viral genomes, a highly reproducible hairpin occurred at about 70 nucleotides from the 5' end of each subunit and contained 320 +/- 8 nucleotides. The stable point of linkage between the subunits in both viral genomes involved fewer than 50 nucleotides and occurred at 466 +/- 9 nucleotides from the 5' end. This places the linkage about 350 nucleotides further toward the 3' end of the subunit than the binding site of primer tRNA. Another structural feature common to both genomes was a loop in each subunit. In M-MuLV, the loop contained 3.9 +/- 0.10 kilobases (kb) and occurred at a distance of 2.2 +/- 0.05 kb from the 5' end. In PR-RSV-A, the loop was smaller (2.3 +/- 0.10 kb) and further (3.3 +/- 0.10 kb) from the 5' end. When M-MuLV RNA was heated to 70, 85, or 90 degrees C and cooled, the hairpin consistently reformed at the 5' end. No other structures typical of the native molecules reappeared. In RNA samples heated to 70 degrees C, a new loop reproducibly occurred near the 5' end of each subunit, but this loop was not found in samples heated to higher temperatures. Based on all of these findings, we conclude that the genome of PR-RSV-A shares several features with M-MuLV and other mammalian retroviruses and that the primer tRNA molecules are not involved in the linkage of the two subunits in either genome. We also conclude that the dimer linkage and the loops in subunits are typical of the native molecules and that their formation requires a special environment.  相似文献   

20.
Carbon-11-labeled serotonin (5-hydroxytryptamine) 6 receptor (5-HT6R) antagonists, 1-[(2-bromophenyl)sulfonyl]-5-[11C]methoxy-3-[(4-methyl-1-piperazinyl)methyl]-1H-indole (O-[11C]2a) and 1-[(2-bromophenyl)sulfonyl]-5-methoxy-3-[(4-[11C]methyl-1-piperazinyl)methyl]-1H-indole (N-[11C]2a), 5-[11C]methoxy-3-((4-methylpiperazin-1-yl)methyl)-1-(phenylsulfonyl)-1H-indole (O-[11C]2b) and 5-methoxy-3-((4-[11C]methylpiperazin-1-yl)methyl)-1-(phenylsulfonyl)-1H-indole (N-[11C]2b), 1-((4-isopropylphenyl)sulfonyl)-5-[11C]methoxy-3-((4-methylpiperazin-1-yl)methyl)-1H-indole (O-[11C]2c) and 1-((4-isopropylphenyl)sulfonyl)-5-methoxy-3-((4-[11C]methylpiperazin-1-yl)methyl)-1H-indole (N-[11C]2c), 1-((4-fluorophenyl)sulfonyl)-5-[11C]methoxy-3-((4-methylpiperazin-1-yl)methyl)-1H-indole (O-[11C]2d) and 1-((4-fluorophenyl)sulfonyl)-5-methoxy-3-((4-[11C]methylpiperazin-1-yl)methyl)-1H-indole (N-[11C]2d), were prepared from their O- or N-desmethylated precursors with [11C]CH3OTf through O- or N-[11C]methylation and isolated by HPLC combined with SPE in 40–50% radiochemical yield, based on [11C]CO2 and decay corrected to end of bombardment (EOB). The radiochemical purity was >99%, and the molar activity (MA) at EOB was 370–740?GBq/μmol with a total synthesis time of ~40-min from EOB.  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