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.
Three new triterpene glucosides, named congmuyenosides C–E ( 1 – 3 , resp.), along with four known ones, were isolated from an EtOH extract of Aralia elata (Miq .) Seem . leaves. The structures of the new compounds were identified as 3‐O‐{β‐D ‐glucopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐(1→3)‐[β‐D ‐glucopyranosyl‐(1→2)]‐β‐D ‐glucopyranosyl}caulophyllogenin ( 1 ), 3‐O‐{β‐D ‐glucopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐(1→3)‐[β‐D ‐glucopyranosyl‐(1→2)]‐β‐D ‐glucopyranosyl}hederagenin 28‐Oβ‐D ‐glucopyranosyl ester ( 2 ), 3‐O‐{β‐D ‐glucopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐(1→3)‐[β‐D ‐glucopyranosyl‐(1→2)]‐β‐D ‐glucopyranosyl}echinocystic acid 28‐Oβ‐D ‐glucopyranosyl ester ( 3 ) on the basis of spectral analyses, including MS, 1H‐NMR, 13C‐NMR, DEPT, HSQC, HMBC, NOESY, and HSQC‐TOCSY experiments. All isolates obtained were evaluated for their cytotoxic activities against three human tumor cell lines (HepG2, SKOV3, and A549). Compound 3 showed significant cytotoxicity against A549 cell line (IC50 9.9±1.5 μM ).  相似文献   

2.
Four different H-type 1 (LedH) blood-group-active glycosphingolipids (LedH-I–IV) have been isolated from the plasma of blood-group O Le(a?b?) secretors. The agglutination of O Le(a?b?) erythrocytes from secretors by 50 μl of 4 hemagglutinating units of caprine anti-LedH (anti-H-type 1) serum was inhibited by 0.02 μg of each of all four glycolipids. No Lea or Leb activities or reaction against Ulex europaeus lectin could be found. LedH-I, -II, -III, and -IV at 0.05, 0.01, 0.01, and 0.02 μg each are sufficient for incubation in order to convert 9 × 107 O Le(a?b?) erythrocytes from nonsecretors into H-type 1 (LedH)-positive cells. Structural analysis of the H-type 1 glycolipids was performed in comparison to that of Lea- and Leb-blood-group-active glycolipids from human plasma isolated previously: Gas chromatography of peracetylated alditols revealed sugar composition. Combined gas chromatography-mass spectrometry established the glycosidic linkages. Together with the results obtained by direct inlet mass spectrometry of permethylated glycosphingolipids and by 360-MHz 1H nuclear magnetic resonance spectroscopy (Egge, H., and Hanfland, P., 1981, Arch. Biochem. Biophys., 210, 396–404; Dabrowski, J., Hanfland, P., Egge, H., and Dabrowski, U., 1981, Arch. Biochem. Biophys., 210, 405–411) the complete structures of the oligosaccharide chains of the Lea-, Leb-, and H-type 1-active glycolipids were established: Galβ1 → 3GlcNAc(4 ← 1αFuc)β1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the Lea antigens; Fucα1 → 2Galβ1 → 3GlcNAc(4 ← 1αFuc)β1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the Leb antigens; and Fucα1 → 2Galβ1 → 3GlcNAcβ1 → 3Galβ1 → 4Glcβ1 → 1 Cer for the H-type 1 (LedH) glycolipids. The diverse antigens of the same blood-group specificity obviously differ from one another in their lipid residue. In addition, plasmatic neolactotetraosylceramide could be identified, differing from that of human erythrocytes by a slower migration behavior in thin-layer chromatography.  相似文献   

3.
Pectins are the most complex polysaccharides of the plant cell wall. Based on the number of methylations, acetylations and glycosidic linkages present in their structures, it is estimated that up to 67 transferase activities are involved in pectin biosynthesis. Pectic galactans constitute a major part of pectin in the form of side‐chains of rhamnogalacturonan‐I. In Arabidopsis, galactan synthase 1 (GALS1) catalyzes the addition of galactose units from UDP‐Gal to growing β‐1,4‐galactan chains. However, the mechanisms for obtaining varying degrees of polymerization remain poorly understood. In this study, we show that AtGALS1 is bifunctional, catalyzing both the transfer of galactose from UDP‐α‐d ‐Gal and the transfer of an arabinopyranose from UDP‐β‐l ‐Arap to galactan chains. The two substrates share a similar structure, but UDP‐α‐d ‐Gal is the preferred substrate, with a 10‐fold higher affinity. Transfer of Arap to galactan prevents further addition of galactose residues, resulting in a lower degree of polymerization. We show that this dual activity occurs both in vitro and in vivo. The herein described bifunctionality of AtGALS1 may suggest that plants can produce the incredible structural diversity of polysaccharides without a dedicated glycosyltransferase for each glycosidic linkage.  相似文献   

4.
The Lea-, Leb-, and H-type 1 (LedH)-blood-group-active glycosphingolipids, as well as H-I-type 2 glycolipid, lactotetraosyl ceramide, and neo-lactotetraosyl ceramide were examined by 1H nuclear magnetic resonance at 360 MHz in dimethyl-d6 sulfoxide as solvent. The resonances of almost all protons of the sugar rings were assigned with the aid of spin decoupling and nuclear Overhauser difference spectroscopy. The latter technique was also applied to establish the sequences and sites of glycosidic linkage. This information, combined with the chemical shift-structure correlations established in our previous work, led to an independent identification of those six glycolipids. Type 1 (Galβ1 → 3GlcNAc) and type 2 (Galβ1 → 4GlcNAc) saccharide chains can be distinguished by this approach. Some deviations from additivity in chemical shifts, calculated for oligosaccharides from the data on their constituent sugar residues, furnished information on the conformational changes in crowded glycolipid molecules.  相似文献   

5.
The interaction of trimannoside, α?benzyl 3, 6‐di‐O‐(α‐D ‐mannopyranosyl)‐α‐D ‐mannopyranoside, 1 with ASAI (Allium sativam agglutinin I, garlic lectin) was studied to reveal the conformational preferences of this ligand in bound‐state and detailed binding mode at atomic level. The binding phenomenon was then compared with another well‐known mannose‐binding lectin, ConA (Concanavalin A). Structural studies of the ligand in free state were done using NMR spectroscopy and Molecular Dynamics simulations. It is found that the substituted‐trimannoside can undergo conformational transitions in solution, with one major and one minor conformation per glycosidic linkage (α 1→3 and α 1→6). On the other hand in the bound‐state only one of the two major conformations was significantly populated. The role of phenyl ring in the binding process was explored. An extended binding site was observed for the trimannoside in ASAI utilizing the aromatic substituent, which is not seen in ConA. Binding data from difference absorption spectroscopy supported this fact that the binding of benzyl‐substituted ligand is tighter with ASAI than ConA. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 952–967, 2010.  相似文献   

6.
The heptasaccharide isolated from the cell wall polysaccharide of Streptococcus mitis J22 serves as an important model for the dynamics and conformation of complex polysaccharides, illustrating the nature of flexibility with rigid epitopes joined by flexible hinges. One-bond C-H residual dipolar couplings (1DCH) and long-range H-H residual dipolar couplings (nDHH) were measured for the heptasaccharide in a cetylpyridinium chloride/hexanol/brine lamellar liquid crystal medium. A method is proposed to determine the nDHH in natural abundance based on a 13C resolved 1H TOCSY pulse sequence previously published to determine the homonuclear scalar couplings. Different methods for interpretation of the 1DCH and the nDHH residual dipolar coupling data obtained were compared and combined with the NOE and long-range H,C and C,C scalar couplings available for this heptasaccharide. A flexible model of the heptasaccharide was determined in which two structurally well-defined regions involving four and two sugar residues, respectively are joined by a flexible hinge which involves two 16 glycosidic linkages.  相似文献   

7.
Residual heteronuclear dipolar couplings obtained from partially oriented protein samples can provide unique NMR constraints for protein structure determination. However, partial orientation of protein samples also causes severe 1 H line broadening resulting from residual 1 H-1H dipolar couplings. In this communication we show that band-selective 1H homonuclear decoupling during data acquisition is an efficient way to suppress residual 1H-1H dipolar couplings, resulting in spectra that are still amenable to solution NMR analysis, even with high degrees of alignment. As an example, we present a novel experiment with improved sensitivity for the measurement of one-bond 1 HN-15N residual dipolar couplings in a protein sample dissolved in magnetically aligned liquid crystalline bicelles.  相似文献   

8.
Two new oleanane‐type saponins: β‐d ‐xylopyranosyl‐(1 → 4)‐6‐deoxy‐α‐l ‐mannopyranosyl‐(1 → 2)‐1‐O‐{(3β)‐28‐oxo‐3‐[(2‐Oβ‐d ‐xylopyranosyl‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐yl}‐β‐d ‐glucopyranose ( 1 ) and 1‐O‐[(3β)‐28‐oxo‐3‐{[β‐d ‐xylopyranosyl‐(1 → 2)‐α‐l ‐arabinopyranosyl‐(1 → 6)‐2‐acetamido‐2‐deoxy‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐yl]β‐d ‐glucopyranose ( 2 ), along with two known saponins: (3β)‐3‐[(β‐d ‐Glucopyranosyl‐(1 → 2)‐β‐d ‐glucopyranosyl)oxy]olean‐12‐en‐28‐oic acid ( 3 ) and (3β)‐3‐{[α‐l ‐arabinopyranosyl‐(1 → 6)‐[β‐d ‐glucopyranosyl‐(1 → 2)]‐β‐d ‐glucopyranosyl]oxy}olean‐12‐en‐28‐oic acid ( 4 ) were isolated from the acetone‐insoluble fraction obtained from the 80% aqueous MeOH extract of Albizia anthelmintica Brongn . leaves. Their structures were identified using different NMR experiments including: 1H‐ and 13C‐NMR, HSQC, HMBC and 1H,1H‐COSY, together with HR‐ESI‐MS/MS, as well as by acid hydrolysis. The four isolated saponins and the fractions of the extract exhibited cytotoxic activity against HepG‐2 and HCT‐116 cell lines. Compound 2 showed the most potent cytotoxic activity among the other tested compounds against the HepG2 cell line with an IC50 value of 3.60μm . Whereas, compound 1 showed the most potent cytotoxic effect with an IC50 value of 4.75μm on HCT‐116 cells.  相似文献   

9.
Introduction – Kudzu root (Radix puerariae) is a rich source of isoflavones that are effective in preventing osteoporosis, heart disease and symptoms associated with menopause. The major isoflavonoids in kudzu root extracts were reported as puerarin, daidzin and daidzein. Recently, an unknown isoflavonoid (compound 1) was detected from one‐year‐old kudzu root cultivated in Vietnam. Objective – To identify a novel compound 1 in kudzu root extract and determine the structure of the compound by ESI+ TOF MS‐MS, 1H‐, 13C‐NMR and enzymatic hydrolysis. Methodology – Samples were prepared by extraction of one‐year‐old kudzu root with 50% ethanol and the isoflavonoids were purified using recycling preparative HPLC. Unknown compound 1 was detected using UV‐light at 254 nm in TLC and HPLC analyses. The molecular weight of 1 was determined using a TOF mass spectrometer equipped with an electrospray ion source. The structure of 1 was determined from the 13C and 1H NMR spectra recorded at 100.40 and 400.0 MHz, respectively. Results – ESI+ TOF MS‐MS analysis shows that 1 is a puerarin diglycoside. The interglycosidic linkage of diglycoside determined by 1H‐, 13C‐NMR, and enzymatic hydrolysis suggests that 1 has a glucosyl residue linked to puerarin by an α‐1,6‐glycosidic bond. This compound is the first naturally‐occurring 8‐[α‐D ‐glucopyranosyl‐(1→6)‐β‐D ‐glucopyranosyl]daidzein in kudzu root. The concentration of glucosyl‐α‐1,6‐puerarin in kudzu root was 2.3 mg/g as determined by HPLC. Conclusion – The results indicate that puerarin diglycoside is one of the major isoflavonoids in kudzu root and has a significant impact on the preparation of highly water‐soluble glycosylated puerarin. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

10.
Lymphocyte function‐associated antigen‐1 (LFA‐1) is an integrin protein that transmits information across the plasma membrane through the so‐called inside‐out and outside‐in signaling mechanisms. To investigate these mechanisms, we carried out an NMR analysis of the dynamics of the LFA‐1 I‐domain, which has enabled us to characterize the motions of this domain on a broad range of timescales. We studied first the internal motions on the nanosecond timescale by spin relaxation measurements and model‐free analysis. We then extended this analysis to the millisecond timescale motions by measuring 15N‐1H residual dipolar couplings of the backbone amide groups. We analyzed these results in the context of the three major conformational states of the I‐domain using their corresponding X‐ray crystallographic structures. Our results highlight the importance of the low‐frequency motions of the LFA‐1 I‐domain in the inactive apo‐state. We found in particular that α‐helix 7 is in a position in the apo‐closed state that cannot be fully described by any of the existing X‐ray structures, as it appears to be in dynamic exchange between different conformations. This type of motion seems to represent an inherent property of the LFA‐1 I‐domain and might be relevant for controlling the access to the allosteric binding pocket, as well as for the downward displacement of α‐helix 7 that is required for the activation of LFA‐1.  相似文献   

11.
A bioassay‐guided phytochemical analysis of the ethanolic extract of Grindelia argentina Deble & Oliveira ‐Deble (Asteraceae) allowed the isolation of a known flavone, hispidulin, and three new oleanane‐type saponins, 3‐Oβ‐D ‐xylopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐2β,3β,16α,23‐tetrahydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester ( 2 ), 3‐Oβ‐D ‐glucopyranosyl‐2β,3β,23‐trihydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester, ( 3 ) and 3‐Oβ‐D ‐xylopyranosyl‐(1→3)‐β‐D ‐glucopyranosyl‐2β,3β,23‐trihydroxyolean‐12‐en‐28‐oic acid 28‐Oβ‐D ‐xylopyranosyl‐(1→2)‐β‐D ‐apiofuranosyl‐(1→3)‐β‐D ‐xylopyranosyl‐(1→3)‐α‐L ‐rhamnopyranosyl‐(1→2)‐α‐L ‐arabinopyranosyl ester ( 4 ), named grindeliosides A–C, respectively. Their structures were determined by extensive 1D‐ and 2D‐NMR experiments along with mass spectrometry and chemical evidence. The isolated compounds were evaluated for their inhibitory activities against LPS/IFN‐γ‐induced NO production in RAW 264.7 macrophages and for their cytotoxic activities against the human leukemic cell line CCRF‐CEM and MRC‐5 lung fibroblasts. Hispidulin markedly reduced LPS/IFN‐γ‐induced NO production (IC50 51.4 μM ), while grindeliosides A–C were found to be cytotoxic, with grindelioside C being the most active against both CCRF‐CEM (IC50 4.2±0.1 μM ) and MRC‐5 (IC50 4.5±0.1 μM ) cell lines.  相似文献   

12.
Saponins are amphiphilic glycoconjugates which give soap‐like foams in H2O. A new triterpenoid saponin, simenoside A ( 1 ), based on gypsogenin aglycone, and the known saponin 2 were isolated from Gypsophila simonii Hub.‐Mor. The structure of the new saponin was elucidated as 3‐Oβ‐D ‐galactopyranosyl‐(1→2)‐[β‐D ‐xylopyranosyl‐(1→3)]‐β‐D ‐glucuronopyranosylgypsogenin 28‐Oβ‐D ‐glucopyranosyl‐(1→3)[β‐D ‐glucopyranosyl‐(1→2)‐β‐D ‐xylopyranosyl‐(1→4)]‐α‐L ‐rhamnopyranosyl‐(1→2)‐β‐D ‐fucopyranosyl ester on the basis of extensive spectral analyses and chemical evidence. Saponins 1 and 2 were isolated from G. simonii for the first time.  相似文献   

13.
Three new steroid saponins (3β,25R)‐spirost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[β‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 1 ), (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐hydroxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 3 ), and (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐hydroxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[β‐D ‐glucopyranosyl‐(1→4)‐6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 5 ), as well as the new pregnane glycoside (3β,16β)‐3‐{[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranosyl]oxy}‐20‐oxopregn‐5‐en‐16‐yl (4R)‐5‐(β‐D ‐glucopyranosyloxy)‐4‐methylpentanoate ( 6 ), were isolated from the rhizomes of Tacca integrifolia together with two known (25R) configurated steroid saponins (3β,25R)‐spirost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 2 ) and (3β,22R,25R)‐26‐(β‐D ‐glucopyranosyloxy)‐22‐methoxyfurost‐5‐en‐3‐yl 6‐deoxy‐α‐L ‐mannopyranosyl‐(1→2)‐[6‐deoxy‐α‐L ‐mannopyranosyl‐(1→3)]‐β‐D ‐glucopyranoside ( 4 ). The cytotoxic activity of the isolated compounds was evaluated in HeLa cells and showed the highest cytotoxicity value for compound 2 with an IC50 of 1.2±0.4 μM . Intriguingly, while compounds 1 – 5 exhibited similar cytotoxic properties between 1.2±0.4 ( 2 ) and 4.0±0.6 μM ( 5 ), only compound 2 showed a significant microtubule‐stabilizing activity in vitro.  相似文献   

14.
Gentiana rhodantha Franch. ex Hemsl. (Gentianaceae), an annual herb widely distributed in the southwest of China, has been medicinally used for the treatment of inflammation, cholecystitis, and tuberculosis by the local people of its growing areas. Chemical investigation on the whole plants led to the identification of eight new phenolic compounds, rhodanthenones A–D ( 1 – 4 , resp.), apigenin 7‐O‐glucopyranosyl‐(1→3)‐glucopyranosyl‐(1→3)‐glucopyranoside ( 5 ), 1,2‐dihydroxy‐4‐methoxybenzene 1‐Oα‐L ‐rhamnopyranosyl‐(1→6)‐β‐D ‐glucopyranoside ( 6 ), 1,2‐dihydroxy‐4,6‐dimethoxybenzene 1‐Oα‐L ‐rhamnopyranosyl‐(1→6)‐β‐D ‐glucopyranoside ( 7 ), and methyl 2‐Oβ‐D ‐glucopyranosyl‐2,4,6‐trihydroxybenzoate ( 8 ), together with eleven known compounds, 9 – 19 . Their structures were determined on the basis of detailed spectroscopic analyses and chemical methods. Acetylcholinesterase (AChE) inhibition and cytotoxicity tests against five human cancer cell lines showed that only rhodanthenone D ( 4 ) and mangiferin ( 12 ) exhibited 18.4 and 13.4% of AChE inhibitory effects at a concentration of 10−4 M , respectively, while compounds 1 – 5 and the known xanthones lancerin ( 11 ), mangiferin ( 12 ), and neomangiferin ( 13 ) displayed no cytotoxicity at a concentration of 40 μM .  相似文献   

15.

Objective:

Galectins (Gal) exert many activities, including regulation of inflammation and adipogenesis. We evaluated modulation of Gal‐1, ‐3, ‐9 and ‐12 in visceral (VAT) and subcutaneous (SAT) adipose tissue in mice.

Design and Methods:

We used two mouse models of obesity, high‐fat diet induced obesity (DIO) and ob/ob mice. We also evaluated the response of Gal‐1 KO mice to DIO.

Results:

Both age and diet modulated expression of galectins, with DIO mice having higher serum Gal‐1 and Gal‐3 versus lean mice after 13‐17 weeks of high‐fat diet. In DIO mice there was a progressive increase in expression of Gal‐1 and Gal‐9 in SAT, whereas Gal‐3 increased in both VAT and SAT. Expression of Gal‐12 declined over time in VAT of DIO mice, similar to adiponectin. Obesity lead to increased production of Gal‐1 in adipocytes, whereas the increased Gal‐3 and Gal‐9 of obesity mostly derived from the stromovascular fraction. Expression of Gal‐12 was restricted to adipocytes. There was increased production of Gal‐3 and Gal‐9, but not Gal‐1, in CD11c? and CD11c+ macrophages from VAT of DIO versus lean mice. Expression of Gal‐1, ‐3 and ‐12 in VAT and SAT of ob/ob mice followed a trend comparable to DIO mice. Rosiglitazone reduced serum Gal‐1, but not Gal‐3 and modulated expression of Gal‐3 in VAT and Gal‐9 and Gal‐12 in SAT of DIO mice. High‐fat feeding lead to increased adiposity in Gal‐1 KO versus WT mice, with loss of correlation between leptin and adiposity and no alterations in glucose and insulin levels.

Conclusions:

Obesity leads to differential modulation of Gal‐1, 3, 9 and 12 in VAT and SAT, with Gal‐1 acting as a modulator of adiposity.
  相似文献   

16.
A hallmark of oligosaccharides is their often limited spatial flexibility, allowing them to access a distinct set of conformers in solution. Viewing each individual or even the complete ensemble of conformations as potential binding partner(s) for lectins in protein–carbohydrate interactions, it is pertinent to address the question on the characteristics of bound state conformation(s) in solution. Also, it is possible that entering the lectin’s binding site distorts the low-energy topology of a glycosidic linkage. As a step to delineate the strategy of ligand selection for galactosides, a common physiological docking point, we have performed a NMR study on two non-homologous lectins showing identical monosaccharide specificity. Thus, the conformation of lactose analogues bound to bovine heart galectin-1 and to mistletoe lectin in solution has been determined by transferred nuclear Overhauser effect measurements. It is demonstrated that the lectins select the syn conformation of lactose and various structural analogues (Galβ(1→4)Xyl, Galβ(1→3)Xyl, Galβ(1→2)Xyl, and Galβ(1→3)Glc) from the ensemble of presented conformations. No evidence for conformational distortion was obtained. Docking of the analogues to the modeled binding sites furnishes explanations, in structural terms, for exclusive recognition of the syn conformer despite the non-homologous design of the binding sites.  相似文献   

17.
The l ‐ascorbate (AsA) content and the expression of six l ‐galactose pathway‐related genes were analyzed in peach flesh during fruit development. Fluctuation of AsA during peach fruit development was divided into four phases based on the overall total AsA (T‐AsA) content per fruit: AsA I, 0–36 days after full bloom (DAFB); AsA II, 37–65 DAFB; AsA III, 66–92 DAFB and AsA IV, 93–112 DAFB. Phase AsA III was a lag phase for AsA accumulation, but did not coincide with the lag phase for fruit development. The T‐AsA concentration was highest at the early stage until 21 DAFB [2–3μmol per gram of fresh weight (g?1 FW)], and decreased to 1/4 and 1/15 of this value at 50 and 92 DAFB, respectively. T‐AsA then remained at 0.15–0.20μmol g?1 FW until harvest at 112 DAFB. More than 90% of the T‐AsA was in the reduced form until 21 DAFB. The proportion of reduced form of AsA then decreased concomitantly with the decrease in AsA concentration. To determine the main pathway of AsA biosynthesis and the AsA biosynthetic capacity of peach flesh, several precursors were incubated with immature whole fruit (59 DAFB). The AsA concentration increased markedly with l ‐galactono‐1,4‐lactone or l ‐galactose (Gal), but d ‐galacturonate and l ‐gulono‐1,4‐lactone failed to increase AsA, indicating dominance of the Gal pathway and potent AsA biosynthetic capabilities in immature peach flesh. The expression of genes involved in the last six steps of the Gal pathway was measured during fruit development. The genes studied included GDP‐d ‐mannose pyrophosphorylase (GMPH), GDP‐ d ‐mannose‐3′,5′‐epimerase (GME), GDP‐ l ‐galactose guanylyltransferase (GGGT), l ‐galactose‐1‐phosphate phosphatase (GPP), l ‐galactose‐1‐dehydrogenase (GDH) and l ‐galactono‐1,4‐lactone dehydrogenase (GLDH). GMPH, GME and GGGT had similar expression patterns that peaked at 43 DAFB. GPP, GDH and GLDH also had similar expression patterns that peaked twice at 21 and 91 DAFB, although the expression of GDH was quite low. High level of T‐AsA concentration was roughly correlated with the level of gene expression in the early period of fruit development (AsA I), whereas no such relationships were apparent in the other periods (e.g. AsA III and IV). On the basis of these findings, we discuss the regulation of AsA biosynthesis in peach fruit.  相似文献   

18.
Two fluorene‐based boronic acids, 9,9‐dimethyl‐9H‐fluoren‐2‐yl‐2‐boronic acid (1) and 9,9‐dimethyl‐9H‐fluoren‐2,7‐diyl‐2,7‐diboronic acid (2), were synthesized and their sensing abilities for detection of D ‐monosaccharides were investigated by fluorescence at physiological pH. It was found that both boronic acids 1 and 2 have high selectivity and sensitivity for D ‐fructose with stability constant of 47.2 and 412.9, respectively. The sensor 2 showed a linear response toward D ‐fructose in the concentration range from 5 × 10–5 to 10–1 mol L–1 with the detection limit of 2 × 10–5 mol L–1. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

19.
Using UDP-d-galactose : 2-acetamido-2-deoxy-d-glucose 4-β-d-galactosyltransferase (EC 2.4.1.22), several methyl β-lactosides have been prepared with 2H- and/or 13C-enrichment at specific sites to facilitate study by 13C (75 MHz) and 1H (600 MHz) n.m.r. spectroscopy. 13C-Chemical shift assignments were verified and the 1H-spectrum of β-lactoside was fully assigned. Sites of enrichment were selected to permit all of the potential three-bond C-C and C-H couplings through the glycosidic bond to be obtained. Replacement of H-3 of the d-glucose residue of methyl β-lactoside with 2H allowed resolution of C-1–H-4′ coupling in the 600-MHz 1H-spectrum. Single or multiple 13C-enrichment at C-1, C-2, C-3, C-1′, C-3′, and/or C-4′ in the disaccharide allowed observation of intra- and inter-residue couplings. 13C-Spin-lattice relaxation-times (T1) are interpreted in terms of molecular motion in solution. The data suggest that methyl β-lactoside has an extended conformation with little rotation about the glycosidic bond. Inter-residue couplings are best explained by tortion angles of φ ~ 40° and ψ ~ 15°, indicating that the conformations of β-lactoside in solution and in the crystal are similar.  相似文献   

20.
A solid state NMR experiment is introduced for probing relatively slow conformational exchange, based on dephasing and refocusing dipolar couplings. The method is closely related to the previously described Centerband-Only Detection of Exchange or CODEX experiment. The use of dipolar couplings for this application is advantageous because their values are known a priori from molecular structures, and their orientations and reorientations relate in a simple way to molecular geometry and motion. Furthermore the use of dipolar couplings in conjunction with selective isotopic enrichment schemes is consistent with selection for unique sites in complex biopolymers. We used this experiment to probe the correlation time for the motion of 13C, 15N enriched urea molecules within their crystalline lattice.  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