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 effect of methylurea, N,N'-dimethylurea, ethylurea, and butylurea as well as guanidine hydrochloride (GuHCl), urea and pH on the thermal stability, structural properties, and preferential solvation changes accompanying the thermal unfolding of ribonuclease A (RNase A) has been investigated by differential scanning calorimetry (DSC), UV, and circular dichroism (CD) spectroscopy. The results show that the thermal stability of RNase A decreases with increasing concentration of denaturants and the size of the hydrophobic group substituted on the urea molecule. From CD measurements in the near- and far-UV range, it has been observed that the tertiary structure of RNase A melts at about 3 degrees C lower temperature than its secondary structure, which means that the hierarchy in structural building blocks exists for RNase A even at conditions at which according to DSC and UV measurements the RNase A unfolding can be interpreted in terms of a two-state approximation. The far-UV CD spectra also show that the final denatured states of RNase A at high temperatures in the presence of different denaturants including 4.5 M GuHCl are similar to each other but different from the one obtained in 4.5 M GuHCl at 25 degrees C. The concentration dependence of the preferential solvation change delta r23, expressed as the number of cosolvent molecules entering or leaving the solvation shell of the protein upon denaturation and calculated from DSC data, shows the same relative denaturation efficiency of alkylureas as other methods.  相似文献   

2.
NMR and CD data have previously shown the formation of the T(4) tetraloop hairpin in aqueous solutions, as well as the possibility of the B-to-Z transition in its stem in high salt concentration conditions. It has been shown that the stem B-to-Z transition in T(4) hairpins leads to S (south)- to N (north)-type conformational changes in the loop sugars, as well as anti to syn orientations in the loop bases. In this article, we have compared by means of UV absorption, CD, Raman, and Fourier transform infrared (FTIR), the thermodynamic and structural properties of the T(4) and A(4) tetraloop hairpins formed in 5'-d(CGCGCG-TTTT-CGCGCG)-3' and 5'-d(CGCGCG-AAAA-CGCGCG)-3', respectively. In presence of 5M NaClO(4), a complete B-to-Z transition of the stems is first proved by CD spectra. UV melting profiles are consistent with a higher thermal stability of the T(4) hairpin compared to the A(4) hairpin. Order-to-disorder transition of both hairpins has also been analyzed by means of Raman spectra recorded as a function of temperature. A clear Z-to-B transition of the stem has been confirmed in the T(4) hairpin, and not in the A(4) hairpin. With a right-handed stem, Raman and FTIR spectra have confirmed the C2'-endo/anti conformation for all the T(4) loop nucleosides. With a left-handed stem, a part of the T(4) loop sugars adopt a N-type (C3'-endo) conformation, and the C3'-endo/syn conformation seems to be the preferred one for the dA residues involved in the A(4) tetraloop.  相似文献   

3.
Neolignan glycosides from Symplocos caudata   总被引:1,自引:0,他引:1  
Huo C  Liang H  Zhao Y  Wang B  Zhang Q 《Phytochemistry》2008,69(3):788-795
A phytochemical investigation of the roots of Symplocos caudata Wall (Symplocaceae) resulted in isolation and characterization of four optical isomers of a neolignan glycoside (1-4), a lignan lactone glycoside (5), a phenylpropanoid glycoside (6), as well as two known compounds (7, 8). Their structures were elucidated as (7S,8S)-threo-7,9,9'-trihydroxy-3,3'-dimethoxy-8-O-4'-neolignan-4-O-beta-d-glucopyranoside (1), (7R,8R)-threo-7,9,9'-trihydroxy-3,3'-dimethoxy-8-O-4'-neolignan-4-O-beta-d-glucopyranoside (2), (7R,8S)-erythro-7,9,9'-trihydroxy-3,3'-dimethoxy-8-O-4'-neolignan-4-O-beta-d-glucopyranoside (3), (7S,8R)-erythro-7,9,9'-trihydroxy-3,3'-dimethoxy-8-O-4'-neolignan-4-O-beta-d-glucopyranoside (4), 8R,8'R-matairesinol-4-O-beta-d-xylopyranosyl-(1-->2)-O-beta-d-glucopyranoside (5), 1-O-[beta-d-xylopyranosyl-(1-->6)-O-beta-d-glucopyranosyl]-2,6-dimethoxy-4-propenyl-phenol (6), matairesinoside (7), and (R)-1-O-(beta-d-glucopyranosyl)-2-[2-methoxy-4-(omega-hydroxypropyl)-phenoxyl]-propan-3-ol (8) on the basis of spectroscopic data (1D and 2D NMR, MS and CD) and chemical evidence.  相似文献   

4.
Nine compounds were isolated from Gentianopsis barbata var. stennocalyx H. W. Li ex. T.N.Ho. Their structures are identified as 1-hydroxy4, 7, 8-trimethoxyxanthone (Ⅰ), 1, 7-dihydroxy-3, 8-dimethoxyxanthone (Ⅱ), 1, 7, 8-trihydroxy-3-methoxyxanthone (Ⅲ), 1-O-(β-D- xylopyranosyl-(1→6)-β-D-glucopyranosyl)-3, 7, 8-trimethoxyxanthone (Ⅳ), 1-O-(β-D-xylopy- ranosyl- (1→6)-β-D-glucopyranosyl)-7-hydroxy-3, 8-dimethoxyxanthone (Ⅴ), 1-O-(β-D-xylo- pyranosyl-(1→6)-β-D-glucopyranosyl)-7, 8-dihydroxy-3-methoxyxanthone (Ⅵ), luteolin-7-O- β-D-glucoside (Ⅶ), oleanolic acid (Ⅷ) and ursolic acid (Ⅸ) by means of chemical methods and UV, IR and NMR determinations respectively.  相似文献   

5.
Cannabinoid receptors are G-protein-coupled receptors comprised of seven transmembrane helices. We hypothesized that the extended helix of the receptor interacts differently with POPC bilayers due to the differing distribution of charged amino acid residues. To test this, hCB1(T377-E416) and hCB2(K278-H316) peptides were studied with 31P and 2H solid-state NMR spectroscopy by incorporating them into 1-palmitoyl-2-oleoyl-sn-glycerophosphocholine bilayers. Lipid affinities of the 40- and 39-residue peptides were analyzed on the basis of 31P and 2H spectral line shapes, order parameters, and T1 relaxation measurements of the POPC bilayers. Lipid headgroup perturbations were noticed in the 31P NMR spectra in the lipid/peptide mixtures when compared with the pure lipids. 2H order parameters were calculated from the quadrupolar splitting of the de-Paked 2H NMR spectra. At the top of the acyl chain, pure lipids had an average S(CD) approximately = 0.20, whereas S(CD) approximately = 0.16 and S(CD) approximately = 0.18 were found in the presence of hCB1(T377-E416) and hCB2(K278-H316), respectively. S(CD) values decreased in the central part of the acyl chains when compared to the pure POPC lipids, indicating a change in the dynamic properties of the lipid membrane in the presence of the cannabinoid peptides. R(1Z) vs S2(CD) plots exhibited a linear dependency with and without the peptides, with an increase in slope upon addition of the peptides to the POPC, indicating that the dynamics of the lipid bilayer is dominated by fast axially symmetric motion. This study provides insights into the interaction of cannabinoid peptides with the membrane bilayer by investigating the headgroup and acyl chain dynamics.  相似文献   

6.
Meneni SR  Shell SM  Gao L  Jurecka P  Lee W  Sponer J  Zou Y  Chiarelli MP  Cho BP 《Biochemistry》2007,46(40):11263-11278
A systematic spectroscopic and computational study was conducted in order to probe the influence of base sequences on stacked (S) versus B-type (B) conformational heterogeneity induced by the major dG adduct derived from the model carcinogen 7-fluoro-2-aminofluorene (FAF). We prepared and characterized eight 12-mer DNA duplexes (-AG*N- series, d[CTTCTAG*NCCTC]; -CG*N- series, d[CTTCTCG*NCCTC]), in which the central guanines (G*) were site-specifically modified with FAF with varying flanking bases (N = G, A, C, T). S/B heterogeneity was examined by CD, UV, and dynamic 19F NMR spectroscopy. All the modified duplexes studied followed a typical dynamic exchange between the S and B conformers in a sequence dependent manner. Specifically, purine bases at the 3'-flanking site promoted the S conformation (G > A > C > T). Simulation analysis showed that the S/B energy barriers were in the 14-16 kcal/mol range. The correlation times (tau = 1/kappa) were found to be in the millisecond range at 20 degrees C. The van der Waals energy force field calculations indicated the importance of the stacking interaction between the carcinogen and neighboring base pairs. Quantum mechanics calculations showed the existence of correlations between the total interaction energies (including electrostatic and solvation effects) and the S/B population ratios. The S/B equilibrium seems to modulate the efficiency of Escherichia coli UvrABC-based nucleotide excision repair in a conformation-specific manner: i.e., greater repair susceptibility for the S over B conformation and for the -AG*N- over the -CG*N- series. The results indicate a novel structure-function relationship, which provides insights into how bulky DNA adducts are accommodated by UvrABC proteins.  相似文献   

7.
The new flavonoid glycoside kaempferol-3-O-alpha-L-rhamnopyranosyl(1-->2)-O-[alpha-L-rhamnopyranosyl(1-->6)]-O-beta-D-galactopyranoside-7-O-alpha-L-rhamnopyranoside was isolated together with (S)-zierin from the leaves of Zollernia ilicifolia (Fabaceae), a medicinal plant used as analgesic and antiulcerogenic effects in Brazilian Tropical Atlantic Rain Forest. The structures were established on the basis of 1H, 13C NMR and 2D NMR (COSY, HMBC, HMQC), UV, MS and IV spectra. The infusion of Zollernia ilicifolia was qualitatively compared to the infusion of the espinheiras-santas (Maytenus aquifolium and Maytenus ilicifolia) by HPLC-DAD.  相似文献   

8.
Eight identical flavonoids were isolated from the alcoholic extract of the pollen of Typha davidiana Hand.-Mazz., Typha latifolia L. and Typha angustata Bory et Chaub. On the basis of spectroscopic (IR, UV, MS and NMR) analysis and the identification of the acidic hydrolytic products, they were identified as naringenin (Ⅰ), isorhamnetin (Ⅱ), quercetin (Ⅲ), isorhamnetin-3-O-(2G-α-L-rhamnopyrano-syl)-rutinoside(Ⅳ), quercetin-3-O-(2G-α-L-rhamnopyranosyl)-rutinoside (Ⅴ), iso-rhamnetin-3-O-rutinoside (Ⅵ),isorhamnetin-3-O-neohesperidoside (Ⅶ), kampferol-3-O-neohesperidoside (Ⅷ). Compound Ⅴ was found in this genus for the first time.  相似文献   

9.
Pericentric inversion of chromosome 9 involving the qh region is relatively common as a constitutional genetic aberration without any apparent phenotypic consequences. However, it has not been established as an acquired abnormality in cancer. Among the three patients reported so far in the literature with acquired inv(9), only one had acute myeloid leukemia (AML). Here we describe an unique case where both chromosomes 9 presented with an acquired pericentric inversion with breakpoints at 9p13 and 9q12 respectively, in a AML patient with aberrant CD7 and CD9 positivity. Additionally, one der(9) also showed short arm deletion at 9p21 to the centromeric region and including the p16 gene. The constitutional karyotype was normal. This is probably the first report describing an acquired inv(9) involving both chromosomes 9 in AML. The possible significance of this inversion is discussed.  相似文献   

10.
Pericentric inversions of the human Y chromosome (inv(Y)) are the result of breakpoints in Yp and Yq. Whether these breakpoints occur recurrently on specific hotspots or appear at different locations along the repeat structure of the human Y chromosome is an open question. Employing FISH for a better definition and refinement of the inversion breakpoints in 9 cases of inv(Y) chromosomes, with seemingly unvarying metacentric appearance after banding analysis, unequivocally resulted in heterogeneity of the pericentric inversions of the human Y chromosome. While in all 9 inv(Y) cases the inversion breakpoints in the short arm fall in a gene-poor region of X-transposed sequences proximal to PAR1 and SRY in Yp11.2, there are clearly 3 different inversion breakpoints in the long arm. Inv(Y)-types I and II are familial cases showing inversion breakpoints that map in Yq11.23 or in Yq11.223, outside the ampliconic fertility gene cluster of DAZ and CDY in AZFc. Inv(Y)-type III shows an inversion breakpoint in Yq11.223 that splits the DAZ and CDY fertility gene-cluster in AZFc. This inversion type is representative of both familial cases and cases with spermatogenetic impairment. In a further familial case of inv(Y), with almost acrocentric morphology, the breakpoints are within the TSPY and RBMY repeat in Yp and within the heterochromatin in Yq. Therefore, the presence of specific inversion breakpoints leading to impaired fertility in certain inv(Y) cases remains an open question.  相似文献   

11.
In temperate lakes, asynchronous cycles in surface water temperatures and incident ultraviolet (UV) radiation expose aquatic organisms to damaging UV radiation at different temperatures. The enzyme systems that repair UV‐induced DNA damage are temperature dependent, and thus potentially less effective at repairing DNA damage at lower temperatures. This hypothesis was tested by examining the levels of UV‐induced DNA damage in the freshwater crustacean Daphnia pulicaria in the presence and absence of longer‐wavelength photoreactivating radiation (PRR) that induces photoenzymatic repair (PER) of DNA damage. By exposing both live and dead (freeze‐killed) Daphnia as well as raw DNA to UV‐B in the presence and absence of PRR, we were able to estimate the relative importance and temperature dependence of PER (light repair), nucleotide excision repair (NER, dark repair), and photoprotection (PP). Total DNA damage increased with increasing temperature. However, the even greater increase in DNA repair rates at higher temperatures led net DNA damage (total DNA damage minus repair) to be greater at lower temperatures. Photoprotection accounted for a much greater proportion of the reduction in DNA damage than did repair. Experiments that looked at survival rates following UV exposure demonstrated that PER increased survival rates. The important implication is that aquatic organisms that depend heavily on DNA repair processes may be less able to survive high UV exposure in low temperature environments. Photoprotection may be more effective under the low temperature, high UV conditions such as are found in early spring or at high elevations.  相似文献   

12.
Lua S  Qin H  Lim L  Shi J  Gupta G  Song J 《PloS one》2011,6(11):e27072
T46I is the second mutation on the hVAPB MSP domain which was recently identified from non-Brazilian kindred to cause a familial amyotrophic lateral sclerosis (ALS). Here using CD, NMR and molecular dynamics (MD) simulations, we characterized the structure, stability, dynamics and binding capacity of the T46I-MSP domain. The results reveal: 1) unlike P56S which we previously showed to completely eliminate the native MSP structure, T46I leads to no significant disruption of the native secondary and tertiary structures, as evidenced from its far-UV CD spectrum, as well as Cα and Cβ NMR chemical shifts. 2) Nevertheless, T46I does result in a reduced thermodynamic stability and loss of the cooperative urea-unfolding transition. As such, the T46I-MSP domain is more prone to aggregation than WT at high protein concentrations and temperatures in vitro, which may become more severe in the crowded cellular environments. 3) T46I only causes a 3-fold affinity reduction to the Nir2 peptide, but a significant elimination of its binding to EphA4. 4) EphA4 and Nir2 peptide appear to have overlapped binding interfaces on the MSP domain, which strongly implies that two signaling networks may have a functional interplay in vivo. 5) As explored by both H/D exchange and MD simulations, the MSP domain is very dynamic, with most loop residues and many residues on secondary structures highly fluctuated or/and exposed to bulk solvent. Although T46I does not alter overall dynamics, it does trigger increased dynamics of several local regions of the MSP domain which are implicated in binding to EphA4 and Nir2 peptide. Our study provides the structural and dynamic understanding of the T46I-causing ALS; and strongly highlights the possibility that the interplay of two signaling networks mediated by the FFAT-containing proteins and Eph receptors may play a key role in ALS pathogenesis.  相似文献   

13.
6-O-[6-O-(N-acetyl-beta-D-glucosaminyl)-N-acetyl-beta-D-glucosaminyl]cyclomaltoheptaose (beta CD) and three positional isomers of 6(1),6(n)-di-O-(N-acetyl-beta-D-glucosaminyl)cyclomaltoheptaose (n=2, 3, and 4) in a mixture of products from beta CD and N-acetylglucosamine by the reversed reaction of beta-N-acetylhexosaminidase from jack bean were isolated and purified by HPLC. The structures of four isomers of di-N-acetylglucosaminyl-beta CDs were determined by FABMS and NMR spectroscopy. The degree of polymerization of the branched oligosaccharides produced by enzymatic degradation with bacterial saccharifying alpha-amylase (BSA) was established by LC-MS methods.  相似文献   

14.
Malonylated flavonol glycosides from the petals of Clitoria ternatea   总被引:2,自引:0,他引:2  
Kazuma K  Noda N  Suzuki M 《Phytochemistry》2003,62(2):229-237
Three flavonol glycosides, kaempferol 3-O-(2"-O-alpha-rhamnosyl-6"-O-malonyl)-beta-glucoside, quercetin 3-O-(2"-O-alpha-rhamnosyl-6"-O-malonyl)-beta-glucoside, and myricetin 3-O-(2",6"-di-O-alpha-rhamnosyl)-beta-glucoside were isolated from the petals of Clitoria ternatea cv. Double Blue, together with eleven known flavonol glycosides. Their structures were identified using UV, MS, and NMR spectroscopy. They were characterized as kaempferol and quercetin 3-(2(G)- rhamnosylrutinoside)s, kaempferol, quercetin, and myricetin 3-neohesperidosides, 3-rutinosides, and 3-glucosides in the same tissue. In addition, the presence of myricetin 3-O-(2"-O-alpha-rhamnosyl-6"-O-malonyl)-beta-glucoside was inferred from LC/MS/MS data for crude petal extracts. The flavonol compounds identified in the petals of C. ternatea differed from those reported in previous studies.  相似文献   

15.
A sizable fraction of T cells expressing the NK cell marker NK1.1 (NKT cells) bear a very conserved TCR, characterized by homologous invariant (inv.) TCR V alpha 24-J alpha Q and V alpha 14-J alpha 18 rearrangements in humans and mice, respectively, and are thus defined as inv. NKT cells. Because human inv. NKT cells recognize mouse CD1d in vitro, we wondered whether a human inv. V alpha 24 TCR could be selected in vivo by mouse ligands presented by CD1d, thereby supporting the development of inv. NKT cells in mice. Therefore, we generated transgenic (Tg) mice expressing the human inv. V alpha 24-J alpha Q TCR chain in all T cells. The expression of the human inv. V alpha 24 TCR in TCR C alpha(-/-) mice indeed rescues the development of inv. NKT cells, which home preferentially to the liver and respond to the CD1d-restricted ligand alpha-galactosylceramide (alpha-GalCer). However, unlike inv. NKT cells from non-Tg mice, the majority of NKT cells in V alpha 24 Tg mice display a double-negative phenotype, as well as a significant increase in TCR V beta 7 and a corresponding decrease in TCR V beta 8.2 use. Despite the forced expression of the human CD1d-restricted TCR in C alpha(-/-) mice, staining with mCD1d-alpha-GalCer tetramers reveals that the absolute numbers of peripheral CD1d-dependent T lymphocytes increase at most by 2-fold. This increase is accounted for mainly by an increased fraction of NK1.1(-) T cells that bind CD1d-alpha-GalCer tetramers. These findings indicate that human inv. V alpha 24 TCR supports the development of CD1d-dependent lymphocytes in mice, and argue for a tight homeostatic control on the total number of inv. NKT cells. Thus, human inv. V alpha 24 TCR-expressing mice are a valuable model to study different aspects of the inv. NKT cell subset.  相似文献   

16.
This work deals with the Dakin-West synthesis, starting from the nucleoamino acid 1-thyminyl acetic acid, as well the NMR, ESI MS, and X-ray characterization of a heteroaromatic compound denominated by us T(2)CO, comprising two thymine moieties anchored to a 2-propanonic unit, the spectroscopic properties of which were studied by UV as a function of temperature and ionic strength. Preliminary binding-studies with molecules of biomedical interest such as nucleic acids and proteins, performed on samples containing T(2)CO, suggested that this molecule is able to interact very weakly with double-stranded RNA, whereas it does not seem to bind other nucleic acids or proteins. Moreover, by studies with fresh human serum we found that T(2)CO is resistant to enzymatic degradation till 24?h, whereas UV metal binding-studies, performed using solutions of copper (II) chloride dihydrate and nickel (II) chloride hexahydrate, revealed a certain ability of T(2)CO to bind copper (II) cation. Finally, by CD spectroscopy we investigated the influence of T(2)CO on the already described supramolecular networks based on L-serine-containing nucleopeptides. More particularly, we found that T(2)CO is able to increase the level of structuration of the non-covalent supramolecular assembly of the chiral nucleopeptides, which is a feature of remarkable interest for the development of innovative drug delivery tools.  相似文献   

17.
The effect of hexafluoroacetone hydrate (HFA) on the structure of the honey bee venom peptide melittin has been investigated. In aqueous solution at low pH melittin is predominantly unstructured. Addition of HFA at pH approximately 2.0 induces a structural transition from the unstructured state to a predominantly helical conformation as suggested by intense diagnostic far UV CD bands. The structural transition is highly cooperative and complete at 3.6 M (50% v/v) HFA. A similar structural transition is also observed in 2,2,2 trifluoroethanol which is complete only at a cosolvent concentration of approximately 8 M. Temperature dependent CD experiments support a 'cold denaturation' of melittin at low concentrations of HFA, suggesting that selective solvation of peptide by HFA is mediated by hydrophobic interactions. NMR studies in 3.6 M HFA establish a well-defined helical structure of melittin at low pH, as suggested by the presence of strong NH/NHi+1 NOEs throughout the sequence, along with many medium range helical NOEs. Structure calculations using NOE-driven distance constraints reveal a well-ordered helical fold with a relatively flexible segment around residues T10-G11-T12. The helical structure of melittin obtained at 3.6 M HFA at low pH is similar to those determined in methanolic solution and perdeuterated dodecylphosphocholine micelles. HFA as a cosolvent facilitates helix formation even in the highly charged C-terminal segment.  相似文献   

18.
Human CD1d molecules present an unknown ligand, mimicked by the synthetic glycosphingolipid alpha-galactosylceramide (alphaGC), to a highly conserved NKT cell subset expressing an invariant TCR Valpha24-JalphaQ paired with Vbeta11 chain (Valpha24(+)Vbeta11(+) invariant NK T cell (NKT(inv))). The developmental pathway of Valpha24(+)Vbeta11(+)NKT(inv) is still unclear, but recent studies in mice were consistent with a TCR instructive, rather than a stochastic, model of differentiation. Using CD1d-alphaGC-tetramers, we demonstrate that in humans, TCR variable domains other than Valpha24 and Vbeta11 can mediate specific recognition of CD1d-alphaGC. In contrast to Valpha24(+)Vbeta11(+)NKT(inv) cells, Valpha24(-)/CD1d-alphaGC-specific T cells express either CD8alphabeta or CD4 molecules, but they are never CD4 CD8 double negative. We show that CD8alphabeta(+)Valpha24(-)/CD1d-alphaGC-specific T cells exhibit CD8-dependent specific cytotoxicity and have lower affinity TCRs than Valpha24(+)/CD1d-alphaGC-specific T cells. In conclusion, our results demonstrate that, contrary to the currently held view, recognition of CD1d-alphaGC complex in humans is not uniformly restricted to the Valpha24-JalphaQ/Vbeta11 NKT cell subset, but can be mediated by a diverse range of Valpha and Vbeta domains. The existence of a diverse repertoire of CD1d-alphaGC-specific T cells in humans strongly supports their Ag-driven selection.  相似文献   

19.
Ultraviolet absorption (UV) and circular dichroism (CD) spectra of wheat germ 5S RNA, when compared to tRNAPhe, indicate a largely base-paired and base-stacked helical structure, containing up to 36 base pairs. Fourier-transform infrared (FT-IR) spectra of tRNAPhe and wheat germ ribosomal 5S RNA have been acquired at 30 and 90 degrees C. From the difference of the FT-IR spectra between 90 and 30 degrees C, the number of base pairs in both RNAs was determined by modification of a previously published procedure [Burkey, K. O., Marshall, A. G., & Alben, J. O. (1983) Biochemistry 22, 4223-4229]. The base-pair composition and total base-pair number from FT-IR data are now consistent for the first time with optical (UV, CD, Raman) and NMR results for ribosomal 5S RNA. Without added Mg2+, tRNAPhe gave 18 +/- 2 base pairs [7 A-U and 11 G-C], in good agreement with the number of secondary base pairs from X-ray crystallography [8 A-U, 12 G-C, and 1 G-U]. Within the 10% precision of the FT-IR method, wheat germ 5S RNA exhibits essentially the same number of base pairs [14 A-U, 17 G-C, and 5 G-U; for a total of 36] in the absence of Mg2+ as in the presence of Mg2+ [14 A-U, 18 G-C, and 3 G-U; for a total of 35], in agreement with the UV hyperchromism estimate of G-C/(A-U + G-C) = 0.58.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The dehydrogenative polymerization of coniferyl alcohol by horseradish peroxidase was performed in 0.10 M phosphate buffer at 27 degrees C. Dehydrogenative polymer (DHP) from coniferyl alcohol was characterized by size exclusion chromatography (SEC) and nuclear magnetic resonance (NMR) spectroscopy. The ratio of 8-O-4':8-5':8-8' linkages was determined by the 1H NMR spectrum of DHP acetate which had good solubility. In "end-wise like" polymerization (the slow addition of hydrogen peroxide), addition of alpha-cyclodextrin to the medium led to DHP with increased 8-O-4' content and a decrease in 8-5' linkages. Under higher pH conditions, DHP with higher 8-O-4' and 8-5' content was obtained in the presence of alpha-cyclodextrin. In the end-wise polymerization (the slow additions of coniferyl alcohol and hydrogen peroxide), using alpha-cyclodextrin also gave DHP with a 8-O-4' richer structure than that prepared in no additive system. The analysis of thioacidolysis products from DHP supported the results of the alpha-cyclodextrin effects on the 8-O-4'-rich structure of DHP. The 8-O-4' structure in DHP prepared in the presence of alpha-cyclodextrin had racemic form as shown by ozonation.  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