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 The self-condensation of 2(3)-O-glycyl esters of adenosine, adenosine-5-(O-methylphosphate) and P1, P2-diadenosine-5-pyrophosphate in 6.2 mM solutions at pH 8.0 and -5°C in the presence of 12.5 mM poly(U) yields approximately 3 times as much diketopiperazine as reactions without poly(U). As the concentration of 2(3)-O-(glycyl)-P1, P2-diadenosine-5-pyrophosphate is decreased from 6.2 mM to 1.5 mM the yield of diketopiperazine in the presence of poly(U) decreases slightly from 6.6% to 5.2%, whereas, in the absence of poly(U) the yield of diketopiperazine decreases substantially from 2.4% to 0.75%. The enhanced yield of diketopiperazine that is attributed to the template action of poly(U) is temperature dependent and is observed only at temperatures below 10°C (5°C to -5°C) for 6.2 mM 2(3)-O-(glycyl)-adenosine-5-(O-methylphosphate) and below 23°C (15°C to -5°C) for 6.2 mM 2(3)-O-(glycyl)-P1, P2-diadenosine-5-pyrophosphate. The absence of a template effect at high temperatures is attributed to the melting of the organized helices. The hydrolysis half-lives at pH 8.0 and -5°C of 2(3)-O-(glycyl)-adenosine, 2(3)-O-(glycyl)-adenosine-5-(O-methylphosphate), 2(3)-O-(glycyl)-P1, P2-diadenosine-5-pyrophosphate, and 5-O-(glycyl)-adenosine in the presence of poly(U) are substantially larger than their half-lives in the absence of poly(U). The condensation of 2(3)-O-(glycyl)-adenosine yields 5% of 5-O-(glycyl)-adenosine in the presence of poly(U) compared to 0.7% in the absence of poly(U).Abbreviations DKP diketopiperazine - (gly)2 glycylglycine - (gly)3 glycylglycylglycine - AppA-gly 2(3)-O-(glycyl)-P1, P2-diadenosine-5-pyrophosphate - MepA-gly 2(3)-O-(glycyl)-adenosine-5-(O-methylphosphate) - Ado-2(3)-gly 2(3)-O-(glycyl)-adenosine - Ado-5-gly 5-O-(glycyl)-adenosine - Boc-gly N-tert-butyloxycarbonylglycine - AppA P1, P2-diadenosine-5-pyrophosphate - MepA adenosine-5-(O-methylphosphate) - AppA-Boc-gly 2(3)-O-(Boc-glycyl)-P1, P2-diadenosine-5-pyrophosphate - Ado-5-Boc-gly 5-O-(Boc-glycyl)-adenosine - Ado-2(3)-Boc-gly 2(3)-O-(Boc-glycyl)-adenosine  相似文献   

2.
Studies of molecular mechanisms of chaperone-like activity of -crystallin became an active field of research over last years. However, fine interactions between -crystallin and the damaged protein and their complex organization remain largely uncovered. Complexation between - and L-crystallins was studied during thermal denaturation of L-crystallin at 60°C using small-angle X-ray scattering (SAXS), light scattering, gel-permeation chromatography, and electrophoresis. A mixed solution of - and L-crystallins at concentrations about 10 mg/ml incubated at 60°C was found to contain their soluble complexes with a mean radius of gyration 14 nm, mean molecular mass 4 MDa and maximal size over 40 nm. In pure L-crystallin solution, no complexes were observed at 60°C. In SAXS studies, transitions in the -crystallin quaternary structure at 60°C were shown to occur and result in doubling of the molecular weight. This suggests that during the temperature-induced denaturation of L-crystallin it binds with modified -crystallin or, alternatively, L-crystallin complexation and -crystallin modifications are concurrent. Estimates of the -L-crystallin complex size and relative contents of - and -L-crystallins in the complex suggest that several -crystallin molecules are involved in complex formation.  相似文献   

3.
Temperature dependencies were established for the egg-to-mummy and mummy-to-adult phases, for mummy mortality, and for parasitism of Aphidius ervi Haliday, Aphidius rhopalosiphi De Stefani-Perez, and Praon volucre (Haliday) (Hymenoptera, Aphidiidae), three parasitoids of Sitobion avenae (Fabricius) (Homoptera, Aphididae), at 8°C, 12°C, 16°C, 20°C, and 25°C on winter wheat (cv. Haven). A physiological model described temperature-dependent development over the full temperature range, whereas a linear model was fitted for data above 8°C and used to estimate the lower temperature thresholds and day-degrees (° D) required for development. The thresholds for A. ervi were 2.2°C for egg-mummy development and 6.6°C for mummy-adult development, those for A. rhopalosiphi were 4.5°C and 7.2°C, and those for P. volucre were 3.8°C and 5.5°C. The time to develop into mummies and adults differed significantly between the three species: A. ervi development into mummies required an average of 159 ° D, while development into adults took an average of 73 ° D. The corresponding average times required for A. rhopalosiphi and P. volucre to develop mummies were 124° D and 126° D, while their development into adults required an average of 70° D and 150° D, respectively. Mummy mortality was 25–35% at 8°C and less at the higher temperatures tested, but began to increase again at 25°C, showing a quadratic relationship between mortality and temperature. Parasitization was very low or, in the case of P. volucre, absent up to 12°C and thereafter increased with increasing temperature. The relationship between parasitization, recorded as percent aphids mummified, and temperature was linear at the temperatures tested and depended on species. A. ervisuperparasitized 11.1% aphids at 20°C and 16.6% aphids at 25°C, whereas superparasitism was low in A. rhopalosiphi and absent in P. volucre. From 16°C to 25°C the P. volucre sex ratio increased. For A. ervi and A. rhopalosiphi there was no trend with temperature, but at 20°C and 25°C it was close to even. Field data for 1996 and 1997 allowed for a comparison of actual and expected emergence of overwintering mummies. In both years, parasitoids were predicted to have emerged from overwintering mummies well in advance of the onset of aphid infestation, and more than a month earlier than the first parasitized aphids were found in winter wheat. Observations from trap plants in other crops supported the predictions of the models. Other factors that can affect biological control by cereal aphid parasitoids are discussed.  相似文献   

4.
Zusammenfassung In Temperaturkabinen wird die Fortpflanzungspotenz von Myzus persicae (Sulzer) (Herkunft Groß-Lüsewitz) bei Dauertemperaturen von 15 bis 30° auf Kohlrüben (Brassica napus subspec. rapifera) untersucht.Eine Populationsanalyse nach Birch (1948) (intrinsic rate of increase) ergab den höchsten Wachstumsfaktor bei Dauertemperaturen zwischen 20 und 23°.Dauertemperaturen >25° führten zu einer starken Minderung der. Fortpflanzungspotenz. 30° ist die obere Grenze der Fortpflanzung der untersuchten Myzus persicae-Population.
The reproductive potential of the peach-potato aphid (origin Gross-Lüsewitz) was studied at temperatures between 15° and 30° in constant temperature chambers. They were cultivated on Swede (Brassica napus spp. rapifera) which stood in Knop's nutrient under gauze cloches in petri dishes. The production of juvenile larvae and the mortality of the mothers was measured daily. The total of all larvae (including those which were dropped) and the total of larvae on the leaf were separately enumerated. The larvae on the leaf were designated as viable larvae. A population analysis using Birch's method showed a maximum value for the growth factor k (difference between birth and mortality rates) of 23° for the total of all larvae, and of 20° for the viable larvae (Fig. 6). The daily relative growth-ratio was at the same temperatures respectively 1.36 and 1.34 (Table IV). Optimum development of M. persicae on swedes occurs thus between 20° and 23°. The percentages of viable larvae which add to the net production of total larvae are 53, 61, 30, and 24 (Table III) at temperatures of 15, 20, 25, and 30° respectively. The average length of a generation was 18.5 days at 15° and less than 13 days at 28 to 30° (Fig. 5). The multiplication rate per generation was 38 at 15°, 48 at 20°, but only 5.5 at 30° (Fig. 4). The time of development from first-stage larva to adult was 12.5 days at 15°, 5 days at 28° and 6 days at 30° (Table VII). The upper limit, where a weak multiplication was still possible, was at 30°. It is concluded that in regions where such limiting temperatures occur during some part of the day, the temperature can be the major regulating factor of the insect populations.
  相似文献   

5.
The polyamides based on 4-amino-1-methylpyrrol-2-carboxylic acid, 4-amino-1-methylimidazole-2-carboxylic acid, and -alanine that stabilize oligonucleotide duplexes consisting of GC pairs through parallel packing in the minor groove were studied. The initial duplex TTGCGCpGCGCAA melts at 28°C; the TTGCGCp[NH(CH2)3COPyImImNH(CH2)3NH(CH3)2][NH(CH2)3COImImPyNH(CH2)3N(CH3)2]GCGCAA duplex (bisphosphoramidate with parallel orientation of ligands, where Py, Im, and are the residues of 1-methyl-4-aminopyrrol-2-carboxylic and 1-methyl-4-aminoimidazole-2-carboxylic acids and -alanine, respectively), at 48°C; and the TTGCGCp[NH(CH2)3COImImPyNH(CH2)3COImImPyNH(CH2)3N(CH3)2]GCGCAA duplex (a hairpin structure with antiparallel orientation), at 56°C.  相似文献   

6.
The continuous control of the maximum position of the dye absorption band (the zero of the derivative dD ()/d of the cell's optical density D ()) in a nematic matrix is demonstrated experimentally, as a result of changing the angle between the optical axis of a planar-oriented sample and the plane of polarization of absorbed light incident normal to the optical axis. The theory proposed describes quantitatively the experimental dependence (). The rotation of the polarizer with given frequency results in the spectral position modulation of the solute band maximum () within (=0°)–(90°)=700 cm–1.  相似文献   

7.
The morphological development and N uptake patterns of spring barley (Hordeum vulgare L.) genotypes of Northern European (Nordic) and Pacific Northwest US (PNW) origin were compared under two diurnally fluctuating root temperature regimes in solution culture. The two regimes, 15/5°C and 9/5°C day maximum/night minimum temperatures, simulated soil temperature differences between tilled vs. heavy-residue, no-till conditions, respectively, observed during early spring in eastern Washington. Previous field experiments indicated that some of the Nordic genotypes accumulated more N and dry matter than the PNW cultivars during early spring under no-till conditions. The objective of this experiment was to determined whether these differences 1) are dependent on the temperature of the rooting environment, and 2) are correlated with genotypic differences in NH4 + and NO3 uptake. Overall, shoot N and dry matter accumulation was reduced by 40% due to lower root temperatures during illumination. Leaf emergence was slowed by 14 to 22%, and tiller production was also inhibited. All genotypes absorbed more ammonium than nitrate from equimolar solutions, and the proportion of total N absorbed as NH4 + was slightly higher in the 9/5°C than the 15/5°C regime. A Finnish genotype, HJA80201, accumulated significantly more shoot N than the PNW cultivars, Clark and Steptoe, and also more than a Swedish cultivar, Pernilla, in the 9/5°C regime. In the 15/5°C regime Steptoe did not differ in shoot N from the Nordic genotypes, while Clark remained significantly lower. These differences were not correlated to relative propensity for N form. Root lengths of the Nordic genotypes were significantly greater than the PNW genotypes grown under the 9/5°C regime, while the root lengths in the warmer root temperture regime were not significantly different among genotypes. Higher root elongation rates under low soil temperature conditions may be an inherent adaptive mechanism of the Nordic genotypes. Overall, the data indicate that lower maximum daytime temperatures of the soil surface layer likely account for a significant portion of the growth reductions and lower N uptake observed in no-till systems.  相似文献   

8.
The relation between the quantum yield of oxygen evolution of open photosystem II reactions centers (p), calculated according to Weis and Berry (1987), and non-photochemical quenching of chlorophyll fluorescence of plants grown at 19°C and 7°C was measured at 19°C and 7°C. The relation was linear when measured at 19°C, but when measured at 7°C a deviation from linearity was observed at high values of non-photochemical quenching. In plants grown at 7°C this deviation occurred at higher values of non-photochemical quenching than in plants grown at 19°C. The deviations at high light intensity and low temperature are ascribed to an increase in an inhibition-related, non-photochemical quenching component (qI).The relation between the quantum yield of excitation capture of open photosystem II reaction centers (exe), calculated according to Genty et al. (1989), and non-photochemical quenching of chlorophyll fluorescence was found to be non-linear and was neither influenced by growth temperature nor by measuring temperature.At high PFD the efficiency of overall steady state electron transport measured by oxygen-evolution, correlated well with the product of q N and the efficiency of excitation capture (exe) but it deviated at low PFD. The deviations at low light intensity are attributed to the different populations of chloroplasts measured by gas exchange and chlorophyll fluorescence and to the light gradient within the leaf.Abbreviations F0 basic fluorescence - F0 basic fluorescence, thylakoid in energized state - Fm maximal fluorescence - Fm maximum fluorescence in energized state - Fs steady state fluorescence - Fv maximal variable fluorescence - PFD photon flux density - PS IIrc Photosystem II reaction center - qF0 quenching of basic fluorescence - qE energy related quenching - qN non-photochemical quenching:-qf-total quenching - qI inhibition-related quenching - qp photochemical quenching - qr quenching due to state transition - Rd dark respiration - p PS II efficiency of excitation capture of open PS IIrc - pe extrapolated minimal value of p - p0 extrapolated maximal value of p - si quantum efficiency of linear electron transport, calculated from gas exchange measurements based on incident light - sf quantum efficiency of linear electron transport, calculated from fluorescence measurements, based on incident measuring light  相似文献   

9.
The C chemical shift tensors of proteins contain information on the backbone conformation. We have determined the magnitude and orientation of the C chemical shift tensors of two peptides with -helical torsion angles: the Ala residue in G*AL (=–65.7°, =–40°), and the Val residue in GG*V (=–81.5°, =–50.7°). The magnitude of the tensors was determined from quasi-static powder patterns recoupled under magic-angle spinning, while the orientation of the tensors was extracted from C–H and C–N dipolar modulated powder patterns. The helical Ala C chemical shift tensor has a span of 36 ppm and an asymmetry parameter of 0.89. Its 11 axis is 116° ± 5° from the C–H bond while the 22 axis is 40° ± 5° from the C–N bond. The Val tensor has an anisotropic span of 25 ppm and an asymmetry parameter of 0.33, both much smaller than the values for -sheet Val found recently (Yao and Hong, 2002). The Val 33 axis is tilted by 115° ± 5° from the C–H bond and 98° ± 5° from the C–N bond. These represent the first completely experimentally determined C chemical shift tensors of helical peptides. Using an icosahedral representation, we compared the experimental chemical shift tensors with quantum chemical calculations and found overall good agreement. These solid-state chemical shift tensors confirm the observation from cross-correlated relaxation experiments that the projection of the C chemical shift tensor onto the C–H bond is much smaller in -helices than in -sheets.  相似文献   

10.
A cell extract of Thermococcus hydrothermalis, grown for 6 h, gave -glucosidase activity at 14.9 U/l, degrading oligosaccharides and maltose. -Amylase, -glucosidase and pullulanase activities were detected at 289 U/l, 13.5 U/l and 30 U/l respectively in the culture medium after 24 h growth of the archaeum. All of three enzymes, characterised by a half-life time of 1 to 5 h at 95°C, degraded both the (14) and (16) linkages of polysaccharides and the (14) linkages of oligosaccharides. © Rapid Science Ltd. 1998  相似文献   

11.
Mesophyll protoplasts were isolated from leaves of three cultivars of Lycopersicon esculentum (L.) Mill., namely Hilda 72, Rutgers and Rentita, and from the wild tomato species Lycopersicon peruvianum (L.) Mill. Protoplasts from L. peruvianum divided and grew actively in a liquid medium according to Zapata et al. (1977), whereas protoplasts from the tomato cultivars Hilda 72 and Rutgers showed comparable rates for cell division only, when the content of major elements in this medium was reduced to one half of the original concentration and when mannitol as osmoticum was replaced by glucose. In Rentita protoplasts no cell division could be observed in about 15 different modifications of the five basic culture media tested. The morphogenetic potential of these tomato cells was assessed by comparing the root and shoot formation of protoplasts and of leaf explants. L. peruvianum exhibited the highest potential. Calli derived from protoplasts regenerated roots on Murastrige-Skoog agar containing 1 M benzylaminopurine (BAP) plus 10 M indole-3-acetic acid (IAA) and 0.1 M BAP plus 1 M IAA. Shoot formation occurred in the combinations of 10 M BAP with 0.1, 1.0, and 10 M IAA. Plantlets regenerated from the L. peruvianum calli could be grown in soil. No shoots or roots were regenerated from calli of Hilda 72 and Rutgers protoplasts in all combinations of BAP and IAA tested in the range from 0.1 M to 100 M, thus indicating the rather low morphogenetic potential of these protoplasts as compared to protoplasts from L. peruvianum leaves.Abbreviations BAP benzylaminopurine - IAA indole-acetic acid - TMV tobacco mosaic virus  相似文献   

12.
A method is described whereby arrays of samples ofClupea pallasi eggs may be stored during their preparation. The high fertilization potential retained by the eggs during short-term storage allows them to be fertilized synchronously when sample preparation is complete. A variation of the dry method of storage retained maximum fertilization potential (80–85%) of the eggs for about 1 hr, and of milt dilution (18 with 17 S sea water), about 7 hr. Following dry storage, eggs fertilized in salinities of 0–45 showed maximum rates of fertilization in salinities of 10–20, and fertilization rates 50% in salinities of 4.5–42. Salinities of fertilization influenced egg diameter, median hatching time, and larval length at hatching in egg samples transferred 21/2 hr after fertilization to an incubation salinity of 17 at 7°C. Fertilization rates were higher (90–95%) for eggs stored in 17 S at 7°C prior to fertilization. Under such wet storage conditions, maximum fertilization pontential was retained for about 2 hr. Highest fertilization rates (95–96%) were obtained for eggs stored and fertilized in salinities of 12–15. For the species and the area of origin considered (British Columbia), wet storage of eggs should result in maximum fertilization when the eggs are stored at 4°C for a period not greater than 2 hr prior to fertilization in the 12–15 S storage medium.Prepared under the auspices of the Canadian-German Scientific and Technical Cooperation Agreement.  相似文献   

13.
Summary The angular dependence of1JC,H in model compounds related to -linked oligosaccharides has been established by FPT INDO quantum chemical calculations. Values calculated for models of (1 1)-, (1 2)-, (1 3)- and (1 4)-linked disaccharides were compared, and the effect of the orientation of HO-2 elucidated. The angular dependence of1JC,H on the torsional angles H and H and the solvent dielectric constant (s) was characterized in the form:1JC,H = A cos2+B cos + C sin2 + D since + E + Fe. The1JC,H values, measured by DEPT methods for C-1-H-1 and C-X-H-X in cellobiose, cyclic trisaccharide and hexopyranoses were used to adjust the calculated angular dependences. Based on the occurrence of the conformers for agarobiose, neoagarobiose, mannobiose and methyl -xylobioside, the thermodynamically averaged <1JC,H > values were calculated. The results obtained (<1JC-1,H-1 > 162.4, <1JC-4, H-4 > 147.6 Hz for methyl -xylobioside; <1JC-1,H-1 > 162.4 and <1JC-4,H-4] > 147.6 Hz for mannobiose; <1JC-1,H-1 > 162.8 Hz for neo agarobiose and <1JC-1,H-1 > 163.2 Hz for agarobiose) agree well with the experimental values of 162.7, 147.5, 160.4, 147.2, 160.9 and 165.7 Hz, respectively.  相似文献   

14.
As the first step in the transfer of barely yellow dwarf virus resistance and salt tolerance from decaploid tall wheatgrass (Thinopyrum ponticum) into hexaploid bread wheat (Triticum aestivum L.), octoploid intergeneric hybrids (2n = 8x = 56) were synthesized by crossing the tall wheatgrass cultivar Alkar with wheat cvs. Fukuhokomugi (Fuko) and Chinese Spring. (Fuko x Alkar) F1 hybrids were studied in detail. The F1 hybrids were perennial and generally resembled the male wheatgrass parent with regard to morphological features and gliadin profile. Most hybrids were euploid with 56 chromosomes and showed high chromosome pairing. On an average, in 6 hybrids 83.6% of the complement showed chiasmatic association, some between wheat and wheatgrass chromosomes. Such a high homoeologous pairing would be obtained if Ph1, the major homoeologous pairing suppressor in wheat, was somehow inactivated. Some of the Fuko x Alkar hybrids had high pollen fertility (18.5–42.0% with a mean of 31.5%) and high seed fertility (3–29 seeds wtih a mean of 12.3 seeds per spike), offering excellent opportunities for their direct backcrossing onto the wheat parent.  相似文献   

15.
Burak  Eugeny S. 《Hydrobiologia》1997,360(1-3):101-107
Life tables of Moina macrocopa (Straus) cultured at seven foodconcentrations (FC) (Scenedesmus sp., 1.49–1490 mg wet weightl-1, 104–106 cellml-1) were investigated for animals of the first generation(nonadapted animals) and for animals of the third generation (adaptedanimals) cultivated at these FC. Adapted animals showed a trophicpreferendum, i.e. a narrow FC-range at which maximal Rovalues were observed in comparison with nonadapted animals. In adaptedanimals, the maximal Ro was 115.3 individual, observed at 20°C and a FC of 74.5 mg wet weight l-1.  相似文献   

16.
Summary Dominant alleles of the Kr1 and Kr2 genes reduce the crossability of hexaploid wheat with many alien species, including rye and Hordeum bulbosum, with Kr1 having the greater effect. However, a cytological study of wheat ovaries fixed 48 h after pollination showed that the wheat genotypes Highbury (kr1, Kr2) and Chinese Spring (Hope 5B) (kr1, kr2) were crossable with Seneca 60 maize, fertilization occurring in 14.4 and 30.7% of embryo sacs respectively. The latter figure was similar to the 29.7% fertilization found in Chinese Spring (kr1, kr2). Most embryo sacs in which fertilization occurred contained an embryo but lacked an endosperm and where an endosperm was formed it was usually highly aberrant. All three wheat x maize combinations were karyotypically unstable and rapidly eliminated maize chromosomes to produce haploid wheat embryos.  相似文献   

17.
Summary Using the existing restriction map and probes from wheat and pea ct-DNA, seven protein genes have been localized in the chloroplast genome of N. tabacum. On the clock-like map, the location of each gene is indicated by its time zone: the 15.2 kD polypeptide of the cytochrome b/f complex at 315, cytochrome f at 430, LS of RuBPCase at 450, both and subunits of ATP synthase at or near 500, proton-translocating subunit of ATP synthase at 820, subunit of ATP synthase at 840 and the 32 kD protein at 930. The genome organization of Nicotiana chloroplast DNA is similar to spinach.  相似文献   

18.
The nonproteinogenic amino acid, cyclopentenylglycine, is found in certain Flacourtiaceae. This compound may be synthesized by two C1-chain elongations of -ketoglutarate via -ketopimelate (C5+2C1) or by condensation of C4 and C3 units (C4+C3), a pathway not involving -ketopimelate. The following experimental design elucidated the biosynthetic pathway: Idesia polycarpa callus cultures were freshly established from leaf petioles; synthetic -[1,2-14C]ketopimelate was added to the medium and cultures were incubated for 3 weeks. After isolation and separation of free amino acids from the tissues, the radioactivity incorporated into cyclopentenylglycine was determined. The results establish -ketopimelate as a precursor for cyclopentenylglycine, thus providing evidence for the C5+2C1 biosynthetic path.  相似文献   

19.
This paper describes a model of a neural visual system of a higher animal, in which the capability of pattern recognition develops adaptively. To produce the adaptability, we adopted self-organizing cells, and with them modeled feature-detecting cells which were discovered by Hubel and Wiesel and whose plasticity was found by Blakemore and Cooper. Combining the self-organizing cells and the learning principle of a Perceptron-type system, we constructed a model of the whole visual system. The model is also equipped with an eye movement control mechanism for gazing, which reduces the number of selforganizing cells required for pattern recognition, thus contributing to their quick self-organization. Computer simulation and an experiment using a hardware simulator showed that self-organizing cells quickly become sensitive to the features often seen and that the resulted system can classify patterns with a rather small number of feature-detecting cells.  相似文献   

20.
The cleavage of adenosine-5-monophosphate (5-AMP) and guanosine-5-monophosphate (5-GMP) by Ce4+ and lanthanide complex of 2-carboxyethylgermanium sesquioxide (Ge-132) in acidic and near neutral conditions was investigated by NMR , HPLC and measuring the liberated inorganic phosphate at 37°C and 50°C. The results showed that 5-GMP and 5-AMP was converted to guanine (G), 5-monophosphate (depurination of 5-GMP), ribose (depurination and dephosphorylation of 5-GMP), phosphate and adenine (A), 5-monophosphate (depurination of 5-AMP), ribose (depurination and dephosphorylation of 5-AMP), phosphate respectively by Ce4+. In presence of lanthanide complexes, 5-GMP and 5-AMP were converted to guanosine (Guo) and phosphate and adenosine (Ado) and phosphate respectively. The mechanism of cleaving 5-GMP and 5-AMP is hydrolytic scission  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