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.
The novel reductive graphene oxide‐based magnetic molecularly imprinted poly(ethylene‐co‐vinyl alcohol) polymers (rGO@m‐MIPs) were successfully synthesized as adsorbents for six kinds of polychlorinated biphenyls (PCBs) in fish samples. rGO@m‐MIPs was prepared by surface molecular imprinting technique. Besides, Fe3O4 nanoparticles (NPs) were employed as magnetic supporters, and rGO@Fe3O4 was in situ synthesis. Different from functional monomer and cross‐linker in traditional molecularly imprinted polymer, here, 3,4‐dichlorobenzidine was employed as dummy molecular and poly(ethylene‐co‐vinyl alcohol) was adopted as the imprinted polymers. After morphology and inner structure of the magnetic adsorbent were characterized, the adsorbent was employed for disperse solid phase extraction toward PCBs and exhibited great selectivity and high adsorption efficiency. This material was verified by determination of PCBs in fish samples combined with gas chromatography‐mass spectrometry (GC‐MS) method. According to the detection, the low detection limits (LODs) of PCBs were 0.0035–0.0070 µg l−1 and spiked recoveries ranged between 79.90 and 94.23%. The prepared adsorbent can be renewable for at least 16 times and expected to be a new material for the enrichment and determination of PCBs from contaminated fish samples. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

2.
It was shown that racemic (±)‐ 2 [1′‐benzyl‐3‐(3‐fluoropropyl)‐3H‐spiro[[2]benzofuran‐1,4′‐piperidine], WMS‐1813 ] represents a promising positron emission tomography (PET) tracer for the investigation of centrally located σ1 receptors. To study the pharmacological activity of the enantiomers of 2 , a preparative HPLC separation of (R)‐2 and (S)‐2 was performed. The absolute configuration of the enantiomers was determined by CD‐spectroscopy together with theoretical calculations of the CD‐spectrum of a model compound. In receptor binding studies with the radioligand [3H]‐(+)‐pentazocine, (S)‐2 was thrice more potent than its (R)‐configured enantiomer (R)‐2 . The metabolic degradation of the more potent (S)‐enantiomer was considerably slower than the metabolism of (R)‐2 . The structures of the main metabolites of both enantiomers were elucidated by determination of the exact mass using an Orbitrap‐LC‐MS system. These experiments showed a stereoselective biotransformation of the enantiomers of 2 . Chirality, 2011. © 2010 Wiley‐Liss, Inc.  相似文献   

3.
H. Sawai  K. Kuroda  J. Seki  H. Ozaki 《Biopolymers》1996,39(2):173-182
Comparative CD studies have been carried out to characterize the properties of 2′–5′ and 3′–5′ oligoriboadenylates and oligoribouridylates from dimer to decamer. The CD band of the 3′–5′ oligoribonucleotides was larger than that of the 2′–5′ oligoribonucleotides and increased with the increase in chain length, while the CD band of the 2′–5′ oligoribonucleotides increased little beyond the dimer level. The CD analysis of the chain length dependency revealed that the 3′–5′ oligoribonucleotides adopt mainly the base-base stacking interaction, while the base-sugar interaction is predominant in the 2′–5′ oligoribonucleotides. The CD intensity of 3′–5′ oligoribonucleotides decreased to a larger extent at elevated temperatures or in the presence of ethanol compared to that of the 2′–5′ counterparts. Mg2+ or Mn2+ ion enhanced the magnitude of the CD of 3′–5′ octariboadenylate, while a small decrease in the CD was observed by the presence of Mg2+ or Mn2+ ion to the 2′–5′ octariboadenylate. The 3′–5′ oligoribonucleotide is likely conformationally flexible and can form helical ordered structure with strong base-base stacking depending on changes in the environment such as temperature, the presence of Mg2+ ion, or hydrophobicity of the solution. © 1996 John Wiley & Sons, Inc.  相似文献   

4.
Seven dinucleoside monophosphates containing 2′-halogeno-2′-deoxypurine nucleoside residue, dAfl-U, dAcl-U, dAbr-U, dAio-U, dGfl-U, and dIfl-C, were chemically synthesized and investigated by 1H-nmr spectroscopy at 300 MHz. The sugar and backbone conformations of these compounds were analyzed by the spectral pattern of furanose proton resonances; and the extents of base-base interaction were estimated from chemical shifts and their temperature-dependent changes of base-proton resonances. It is found that the population of C3′-endo conformer and the extent of base-base interaction decrease as the electronegativity of 2′-substituent decreases in dAx-U (x = fl, cl, br, and io) series. The C3′-endo (3E) population and the base-base interaction in Nfl-U (N = A,G)-type dimers as well as dIfl-C are relatively higher than the corresponding natural ribo-dimers but can be recognized as grossly similar to the conformation of regular RNA dimers.  相似文献   

5.
Polarized Raman scattering measurements have been made of a single crystal of uridylyl(3′–5′)adenosine (UpA) by the use of a Raman microscope with 488.0 nm excitation. The UpA crystal belongs to space group P21 (monoclinic), and Raman intensities Iaa, Ibb, and Ic′c′, have been determined for each Raman band. These intensities correspond to the aa, bb, and c′c′ components of the crystal Raman tensor, where c′ is defined as an axis perpendicular to the crystallographic a axis in the ac plane. From these experimental data, and by taking the known crystal structure into account, anisotropic and isotropic molecular Raman tensors have been calculated for the following 11 normal modes: ring stretching modes of the adenine residue (protonated) at 1560, 1516, 1330, and 715 cm−1; ring stretching modes of the uracil residue at 1696, 1657, 1615, 1228, and 790 cm−1; PO2 symmetric stretching mode at 1080 cm−1; P(—)O single bond stretching mode at 801 cm−1. These pieces of information of the Raman tensors are considered to be useful for estimating the orientations of the DNA and RNA strands in a biological complex from a polarized Raman spectroscopic measurement of such a complex. © 1998 John Wiley & Sons, Inc. Biopoly 45: 135–147, 1998  相似文献   

6.
Five known secondary metabolites, chrysophanol ( 1 ), 7,7′‐biphyscion ( 2 ), secalonic acid D ( 3 ), mannitol ( 4 ) and trehalose ( 5 ) were isolated for the first time from the extracts of the fungus Phialomyces macrosporus. Their structures were elucidated by NMR methods (1D and 2D NMR analysis), optical activity and ESI‐MS. Complete 1H and 13C assignments were performed for compound 2 . The antimicrobial activity was evaluated by serial microdilution assay for compounds 2 and 3 and results showed that compound 3 exhibited a significant growth inhibition at concentrations of 15.6 mg/ml (S. aureus and S. choleraesius) and 0.97 mg/mL (B. subtilis), comparable to the positive control.  相似文献   

7.
In order to obtain information about the conformational features of a 2′-O-methylated polyribonucleotide at the nearest neighbor level, a detailed nuclear magnetic resonance study of AmpA was undertaken. AmpA was isolated from alkali hydrolysates of yeast RNA, and proton spectra were recorded at 100 MHz in the Fourier transform mode in D2O solutions, 0.01 M, pH 5.4 and 1.5 at 25°C. 31P spectra were recorded at 40.48 MHz. Complete, accurate sets of nmr parameters derived for each nucleotidyl unit by simulation iteration methods. The nmr data were translated into conformational parameters for all the bonds using procedures developed in earlier studies from these laboratories. It is shown that AmpA exists in aqueous solution with a flexible molecular framework, which shows preferences for certain orientations. The ribose rings exist as a 2E ? 3E equilibrium with the —pA ribose showing a bias for the 3E pucker. The C(4′)—C(5′) bonds of both nucleotidyl units show significant preference (75–80%) to exist in gg conformation. The dominant conformer (80%) about C(5′)—O(5′) of the 5′-nucleotidyl unit is gg′. Even though an unambiguous determination of the orientation of the 3′-phosphate group cannot be made, tentative evidence shows that it preferentially occupies g+ domains [O(3′)—P trans to C(3′)—C(2′)] in which the H(3′) —C(3′)—O(3′)—P(3′) dihedral angle is about 31°. There is reasonable evidence that the 2′-O-methyl preferentially occupies the domain in which the O(2′)—CH3 bond is trans to C(2′)—C(1′). Lowering of pH to 1.5, which results in protonation of both the adenine moieties, causes destacking of AmpA. Such destacking is accompanied by small, but real, perturbations in the conformations about most of the bonds in the backbone. A detailed comparison of the solution conformations of ApA and AmpA clearly shows that 2′-O-methylation strongly influences the conformational preference about the C(3′)—O(3′) bond of the 3′-nucleotidyl unit, in addition to inducing small changes in the overall ribophosphate backbone conformational equilibria. The effect of 2′-O-methylation is such that the C(3′)—O(3′) is forced to occupy preferentially the g+ domain rather than the normally preferred g? domain [O(3′)—P trans to C(3′)—C(4′)] in ApA. The data on ApA and AmpA further reveal that the extent of stacking interaction is less in AmpA compared to ApA. It is suggested that stacked species of AmpA exist as right-handed stacks where the magnitude of ω and ω′ about O(5′)—P and P—O(3′) is about 290°. The reason for the lesser degree of stacking in AmpA compared to ApA is intramolecular interaction between 2′-O-methyl and the flexible O(3′)—P—O(5′) bridge, the interaction causing some perturbation in the magnitudes of ω/ω′, causing destacking. The destacking will lead to an increase in χCN by a few degrees, causing an increase in 2E populations; the latter in turn will shift the 3′ phosphate group from g? to g+ domains. In short, a coupled series of conformational events is envisioned at the onset of destacking, made feasible by the interaction between the 2′-O-methyl group and the swivel O(3′)—P—O(5′) bridge.  相似文献   

8.
A series of new acetohydrazone‐containing 1,2,4‐triazolo[1,5‐a]pyrimidine derivatives were designed and synthesized for the purpose of searching for novel agrochemicals with higher fungicidal activity. Their in vitro fungicidal activities against Rhizoctonia solani were evaluated, and the most promising compound, 2‐[(5,7‐dimethyl[1,2,4]triazolo[1,5‐a]pyrimidin‐2‐yl)sulfanyl]‐2′‐[(2‐hydroxyphenyl)methylidene]acetohydrazide ( 2‐17 ), showed a lower EC50 value (5.34 μg ml?1) than that of commercial carbendazim (EC50=7.62 μg ml?1). Additionally, compound 2‐17 was also found to display broad‐spectrum fungicidal activities, and its EC50 value (4.56 μg ml?1) against Botrytis cinereapers was very similar to that of carbendazim. Qualitative structure–activity relationships (QSARs) of the synthesized compounds were also discussed.  相似文献   

9.
Protected dinucleoside‐2′,5′‐monophosphate has been prepared to develop a prodrug strategy for 2‐5A. The removal of enzymatically and thermally labile 4‐(acetylthio)‐2‐(ethoxycarbonyl)‐3‐oxo‐2‐methylbutyl phosphate protecting group and enzymatically labile 3′‐O‐pivaloyloxymethyl group was followed at pH 7.5 and 37 °C by HPLC from the fully protected dimeric adenosine‐2′,5′‐monophosphate 1 used as a model compound for 2‐5A. The desired unprotected 2′,3′‐O‐isopropylideneadenosine‐2′,5′‐monophosphate ( 9 ) was observed to accumulate as a major product. Neither the competitive isomerization of 2′,5′‐ to a 3′,5′‐linkage nor the P–O5′ bond cleavage was detected. The phosphate protecting group was removed faster than the 3′‐O‐protection and, hence, the attack of the neighbouring 3′‐OH on phosphotriester moiety did not take place.  相似文献   

10.
A comparative study has been made of the configurational effects on the conformational properties of α- and β-anomers of purine and pyrimidine nucleoside 3′,5′,-cyclic monophosphates and their 2′-arabino epimers. Correlation between orientation of the base and the 2′-hydroxyl group have been studied theoretically using the PCILO (Perturbative Configuration Interaction using Localized Orbitals) method. The effect of change in ribose puckering on the base-hydroxyl interaction has also been studied. The result show that steric repulsions and stabilizing effects of intramolecular hydrogen bonding between the base and the 2′-hydroxyl (OH) group are of major importance in determining configurations of α-anomers and 2′-arabino-β-epimers. For example, hydrogen bonding between the 2′-hydroxyl group and polar centers on the base ring is clearly implicated as a determinant of syn-anti preferences of the purine (adenine) or pyrimidine (uracil) bases in α-nucleoside 3′,5′-cyclic monophosphates. Moreover, barrier heights for interconversion between conformers are sensitive to ribose pucker and 2′-OH orientations. The result clearly show that a change in ribose-ring pucker plays an essential role in relieving repulsive interaction between the base and the 2′-hydroxyl group. Thus a C2′-exo-C3′-endo (2T3) pucker is favored for α-anomers in contrast with the C4′-exo-C3′-endo (4T3) from found in β-compounds.  相似文献   

11.
In order to obtain information about the conformational characteristics at the nearestneighbor level in the 2′-O-methylated region of t-RNA, as well as in the bizarre 5′-terminus of eucaryotic mRNA, a detailed nuclear magnetic resonance study of 2′-O-methyl-cytidylyl-(3′ → 5′)-cytidine (CmpC) was conducted. Proton spectra were recorded at 270 MHz in the Fourier mode in D2O solutions, 0.01M, pD 7.3 in the temperature range 5–80°C. Complete accurate sets of nmr parameters were derived for each of the nucleotidyl units by a combination of homo-nuclear decouplings and simulation iteration methods. The data were translated into conformational parameters using procedures developed in earlier studies from these laboratories. It is shown that the ribofuranose ring exists at a 2E ? 3E equilibrium with clear preference [(75–80)%] for the 3E mode. The C(4′)-C(5′) and C(5′)-O(5′) bonds form a stable conformational network with outspoken preference for conformers in which Ψ1, Ψ2 ? 60° and ?2 ? 180°. The orientation of the 3′-phosphate and 2′-O-methyl groups is such that ?1′ ? 210° and ?″ ? 60°. The phosphodiester bonds are flexible and shift trends for base, H(1′), and H(5″) suggest the existence of a conformational blend of right-handed stack (g?g?), left-handed stack (g+g+), and unstacked arrays (tg? and tg+). Elevation of temperature perturbs the 2E ? 3E equilibrium accompanied with modest depopulation of ψ1, ψ2 ? 60° and ?2 ? 180° conformers. The major effect of elevation of temperature is in the increase of unstacked arrays at the expense of g?g? and g+g+ conformers. The shift trend of Cmp-H(3′) with temperature shows that torsional variation about O(3′)-P is facilitated by increase in temperature and the preferred rotamer about O(3′)-P in the unstacked form is t (ω1′ = 180°). A detailed comparison of the aqueous solution conformations of CpC and CmpC reveals that 2′-O-methylation causes: (i) a reduction in the magnitude of χ1; (ii) an increase in the population of 3E pucker at the 3′-nucleotidyl unit; and (iii) modest perturbations in the O(3′)-P and P-O(5′) bond conformations. Comparison of the aqueous solution conformations of AmpA and CmpC makes clear that the conformational properties of pyrimidine-pyrimidine and purine-purine dimers which carry a 2′-O-methylated 3′-nucleotidyl unit are significantly different.  相似文献   

12.
13.
In order to analyze the melting behavior of 5′-rGMP gel at acidic pH and self-aggregate near neutral pH we have obtained Raman spectra of aqueous solutions of 5′-rGMP at various temperatures. At low temperature the intensities of Raman peaks at 502, 585, 1083, 1179, 1322, 1366, 1487, and 1578 cm?1 decrease due to the formation of ordered structure (Raman hypochromism). In contrast, the peaks at 671, 725, 813, and 1338 cm?1 become stronger at low temperature (Raman hyperchromism). The Raman hyperchromism of the 671- and 813-cm?1 peaks have been explained in terms of detailed structural models. Recently, the 668- and 682-cm?1 peaks in the Raman spectrum of aqueous 5′-rGMP solution have been attributed to the guanine ring breathing vibrations in C3′- and C2′-endo conformers [Benevides, J. B., Lemur, D. & Thomas, G. J., Jr. (1984) Biopolymers 23 , 1011–1024]. On the basis of this information our Raman data can be interpreted to suggest that the continuous helix model of 5′-rGMP gel is right-handed. The 1487-cm?1 peak intensity has been used to monitor the melting profies at several pHs. Near neutral pH the melting profile shows a single transition, whereas at acidic pH it shows two transitions. From these observations we propose possible pathways for the melting of 5′-rGMP gel formed at acidic pH and self-aggregate formed near neutral pH.  相似文献   

14.
Metoprolol is available for clinical use as a racemic mixture. The S‐(?)‐metoprolol enantiomer is the one expressing higher activity in the blockade of the β1‐adrenergic receptor. The α‐hydroxymetoprolol metabolite also has activity in the blockade of the β1‐adrenergic receptor. The present study describes the development and validation of a stereoselective method for sequential analysis of metoprolol and of α‐hydroxymetoprolol in plasma using high‐performance liquid chromatography with tandem mass spectrometry (LC‐MS/MS). 1‐ml aliquots of plasma were extracted with dichloromethane : diisopropyl ether (1:1, v/v). Metoprolol enantiomers and α‐hydroxymetoprolol isomers were separated on a Chiralpak AD column (Daicel Chemical Industries, New York, NY, USA) and quantitated by LC‐MS/MS. The limit of quantitation obtained was 0.2 ng of each metoprolol enantiomer/ml plasma and 0.1 ng/ml of each α‐hydroxymetoprolol isomer/ml plasma. The method was applied to the study of kinetic disposition of metoprolol in plasma samples collected up to 24 h after the administration of a single oral dose of 100‐mg metoprolol tartrate to a hypertensive parturient with a gestational age of 42 weeks. The clinical study showed that the metoprolol pharmakokinetics is enantioselective, with the observation of higher area under the curve (AUC)0?∞ values for S‐(?)‐metoprolol (AUCS‐(?)/AUCR‐(+) = 1.81) and the favoring of the formation of the new chiral center 1′R of α‐hydroxymetoprolol (AUC0?∞1′R/1′S = 2.78). Chirality, 25:1–7, 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

15.
A simple, sensitive and rapid HPLC method with fluorescence detection for the determination of dimethyl‐4,4′‐dimethoxy‐5,6,5′,6′‐dimethylene dioxybiphenyl‐2,2′‐dicarboxylate (DDB) in the raw material and pill form was developed. Liquid chromatography was performed on a C18 column (250 × 4.6 mm i.d., 5 µm particle size), the mobile phase consisted of methanol and 0.05 M sodium dihydrogen phosphate buffer (80 : 20, v/v), and the apparent pH of the mobile phase was adjusted to 3. The fluorescence detector was operated at excitation/emission wavelengths of 275/400 nm. The proposed method allows the determination of DDB within concentration range 0.1–1.5 µg/mL with a limit of detection of 0.032 µg/mL, a limit of quantification of 0.097 µg/mL and a correlation coefficient of 0.9997. The proposed method has been successfully applied for the analysis of DDB in its pills with a percentage recovery of 98.45 ± 0.32. The method was fully validated according to ICH guidelines. Moreover, the high sensitivity of the method permits its use in an in vitro dissolution test for DDB under simulated intestinal conditions. In addition, the proposed method was extended to a content uniformity test according to USP guidelines. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

16.
The Raman spectra of guanylyl (3′-5′) guanosine (GpG) in solution in H2O and D2O at pH 3–7 have been recorded at various temperatures between 0 and 80°C. The results are consistent with the existence in the lower temperature range of stable aggregates formed by the stacking of GpG tetramers. The aggregates melt cooperatively near 60°C, which results in important changes in the spectra. Among these, a large increase in intensity of some of the bands assigned to the guanine residues shows that unstacking of the bases occurs at the melting. Also apparent in the spectra are changes in the intensity and frequency of band attributable to molecular groups involved in intermolecular hydrogen bonding between adjacent molecules in the complex. The melting temperature of GpG decreases by approximately 15°C upon lowering the concentration from 5 × 10?2 to 5 × 10?4M, as shown by Raman, calorimetric, CD, and uv measurements. The experimentally determined ΔH and ΔS for the melting transition are 9 Kcal/mol and 28 e.u./mol, respectively. The aggregation of GpG in 1.5 × 10?3M solutions was found to be very slow. The half-time of the process, which roughly follows first-order kinetics, is approximately 3 min at 10°C and 21 min at 35°C. The negative energy of activation associated with this reaction (?143 Kcal) indicated that the process involves intermediates whose concentrations decrease the temperatures raised, thus slowing down the overall process. The rate of disaggregation of GpG upon dilution to very low concentration is also extremely slow, indicating that the GpG aggregates, once formed, are very stable.  相似文献   

17.
Penciclovir is a potent antiherpesvirus agent which is highly selective due to its phosphorylation only in virus infected cells. Phosphorylation of one of the hydroxymethyl groups of penciclovir (PCV) creates a chiral centre leading to the possible formation of (R)- and (S)-enantiomers. The absolute configuration and stereospecificity of the PCV-phosphates produced in cells infected with herpes simplex viruses types 1 and 2 (HSV-1 and HSV-2), as well as by HSV-1-encoded thymidine kinase, were determined using isotopically chiral [4′-13C]PCV precursors and 13C NMR spectroscopy of the isolated metabolites. The absolute configuration of penciclovir-triphosphate (PCV-TP) produced in HSV-1-infected cells was shown to be S with an enantiomeric purity of greater than 95%. However, in contrast to HSV-1-infected cells in which none of the (R) enantiomer was detected, about 10% of (R)-PCV-TP was produced in HSV-2-infected cells. Phosphorylation of PCV by HSV-1-encoded thymidine kinase was found to give 75% (S)- and 25% (R)-PCV-monophosphate. The proportion of the (S)-isomer appears to be amplified in the subsequent phosphorylations leading to the triphosphate. © 1993 Wiley-Liss, Inc.  相似文献   

18.
Nucleic acids analogues, i.e., oligonucleotide N3′→P5′ phosphoramidates and N3′→P5′ thio‐phosphoramidates, containing 3′‐amino‐3′‐deoxy nucleosides with various 2′‐substituents were synthesized and extensively studied. These compounds resist nuclease hydrolysis and form stable duplexes with complementary native phosphodiester DNA and, particularly, RNA strands. An increase in duplexes' melting temperature, ΔTm, relative to their phosphodiester counterparts, reaches 2.2–4.0° per modified nucleoside. 2′‐OH‐ (RNA‐like), 2′‐O‐Me‐, and 2′‐ribo‐F‐nucleoside substitutions result in the highest degree of duplex stabilization. Moreover, under close to physiological salt and pH conditions, the 2′‐deoxy‐ and 2′‐fluoro‐phosphoramidate compounds form extremely stable triple‐stranded complexes with either single‐ or double‐stranded phosphodiester DNA oligonucleotides. Melting temperature, Tm, of these triplexes exceeds Tm values for the isosequential phosphodiester counterparts by up to 35°. 2′‐Deoxy‐N3′→P5′ phosphoramidates adopt RNA‐like C3′‐endo or N‐type nucleoside sugar‐ring conformations and hence can be used as stable RNA mimetics. Duplexes formed by 2′‐deoxy phosphoramidates with complementary RNA strands are not substrates for RNase H‐mediated cleavage in vitro. Oligonucleotide phosphoramidates and especially thio‐phosphoramidates conjugated with lipid groups are cell‐permeable and demonstrate high biological target specific activity in vitro. In vivo, these compounds show good bioavailability and efficient biodistribution to all major organs, while exerting acceptable toxicity at therapeutically relevant doses. Short oligonucleotide N3′→P5′ thio‐phosphoramidate conjugated to 5′‐palmitoyl group, designated as GRN163L (Imetelstat), was recently introduced as a potent human telomerase inhibitor. GRN163L is not an antisense agent; it is a direct competitive inhibitor of human telomerase, which directly binds to the active site of the enzyme and thus inhibits its activity. This compound is currently in multiple Phase‐I and Phase‐I/II clinical trials as potential broad‐spectrum anticancer agent.  相似文献   

19.
Potential energy calculations were employed to examine the effect of ribose 2′-O-methylation on the conformation of GpC. Minimum energy conformations and allowed conformational regions were calculated for 2′MeGpC and Gp2′MeC. The two lowest energy conformations of 2′MeGpC and Gp2′MeC are similar to those of GpC itself. The helical RNA conformation (sugar pucker-C(3′)-endo, ω′ and ω,g?g?, bases-anti) is the global minimum, and a helix-reversing conformation with ω′, ω in the vicinity of 20°, 80° is next in energy. However, subtle differences between the three molecules are noted. When the substitution is on the 5′ ribose (Gp2′MeC), the energy of the helical conformation is less than that of GpC, due to favorable interactions of the added methyl group. When the substitution is at the 3′ ribose (2′MeGpC) these stabilizing interactions are outweighed by steric restrictions, and the helical conformation is of higher energy than for GpC. Furthermore, the statistical weight of the 2′MeGpC g? g? helical region is substantially less than the corresponding weight for Gp2′MeC. In addition, 2′MeGpC′s methoxy group is conformationally restricted to a narrow range centered at 76°. This group has a broadly allowed region between 50 and 175° in Gp2′MeC. These differences occur because the appended methyl group in 2′MeGpC is located in the interior of the helix cylinder, as it would be in polynucleotide, while it hangs unimpeded in Gp2′MeC. These findings suggest that 2′-O-methylation has both stabilizing and destabilizing influences on the helical conformation of RNA. For 2′MeGpC the destabilizing steric hindrance imposed by the nature of the guanine base dominates.  相似文献   

20.
Herein, a highly selective high‐performance liquid chromatography (HPLC) coupled with resonance Rayleigh scattering (RRS) method was developed to detect gatifloxacin (GFLX) and sparfloxacin (SPLX). GFLX and SPLX were first separated by HPLC, then, in pH 4.4 Britton–Robinson (BR) buffer medium, protonic quaternary ammonia cation of GFLX and SPLX reacted with erythrosine (ERY) to form 1:1 ion‐association complexes, which resulted in a significant enhancement of RRS signal. The experimental conditions of HPLC and post‐column RRS have been investigated, including detection wavelength, flow rate, pH, reacting tube length and reaction temperature. Reaction mechanism were studied in detail by calculating the distribution fraction. The maximum RRS signals for GFLX and SPLX were recorded at λex = λem = 330 nm. The detection limits were 3.8 ng ml?1 for GFLX and 17.5 ng ml?1 for SPLX at a signal‐to‐noise ratio of 3. The developed method was successfully applied to the determination of GFLX and SPLX in water samples. Recoveries from spiked water samples were 97.56–98.85%.  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