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.
The catalytic effect of individual conventional anionic surfactant, namely, sodium lauryl sulfate (NaLS), anionic gemini surfactant, namely, sodium salt of bis(1‐dodecenyl succinamic acid) (NaBDS), and mixed surfactant (NaLS + NaBDS) on the rate of oxidation of D ‐fructose by alkaline chloramine‐T has been investigated. The reaction always showed a first‐order dependence of rate with respect to each fructose, alkali, and chloramine‐T. The rate was proportional to (k′+k″ [surfactant]), where k′ and k″ are the rate constants in the absence and presence of the surfactant, respectively. The binding parameters have been evaluated. The observed catalytic effect of mixed micelle on the rate of oxidation was always less than the algebraic sum of the catalytic effect of two surfactants when they were taken separately, suggesting an antagonism (negative synergism) in mixed micelle. The antagonism has also been confirmed by determining critical micelle concentration and interaction parameter (βm) of mixed micelle under the experimental conditions of kinetics, that is, in alkaline medium. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 123–132, 2009  相似文献   

2.
Three new phenylethanoid glycosides, named digicilisides A – C ( 1  –  3 , resp.), have been isolated from the roots of Digitalis ciliata, along with five known phenylethanoid glycosides. The structures of 1  –  3 were identified as 2‐(4‐hydroxy‐3‐methoxyphenyl)ethyl β‐d ‐glucopyranosyl‐(1→3)‐[α‐l ‐rhamnopyranosyl‐(1→6)]‐4‐O‐[(E)‐feruloyl]‐β‐d ‐glucopyranoside ( 1 ), 2‐(3,4‐dihydroxyphenyl)ethyl α‐l ‐arabinopyranosyl‐(1→2)‐[β‐d ‐glucopyranosyl‐(1→3)]‐[α‐l ‐rhamnopyranosyl‐(1→6)]‐4‐O‐[(E)‐feruloyl]‐β‐d ‐glucopyranoside ( 2 ), and 2‐(3,4‐dihydroxyphenyl)ethyl β‐d ‐glucopyranosyl‐(1→3)‐{6‐O‐[(E)‐feruloyl]‐β‐d ‐glucopyranosyl‐(1→6)}‐4‐O‐[(E)‐caffeoyl]‐β‐d ‐glucopyranoside ( 3 ).  相似文献   

3.
Four new 5β,19‐epoxycucurbitacins, kuguacins T–W ( 1 – 4 , resp.), along with nine known cucurbitane derivatives, 5 – 13 , were obtained from the fresh fruit of Momordica chrantia. Structures of the new metabolites were elucidated as 5β,19‐epoxy‐25‐hydroxycucurbitane‐3,7,23‐trione ( 1 ), 5β,19‐epoxy‐3,7‐dioxo‐23,24,25,26,27‐pentanorcucurbitan‐22‐oic acid ( 2 ), 5β,19‐epoxy‐3β‐hydroxycucurbit‐24‐ene‐7,23‐dione ( 3 ), and 5β,19‐epoxy‐25‐hydroxycucurbit‐23‐ene‐3,7‐dione ( 4 ), by extensive spectroscopic investigations, which were confirmed by a single‐crystal X‐ray diffraction analyses in the case of compound 4 .  相似文献   

4.
A single boronic acid‐based fluorescent probe (compound CSP) for saccharides was designed and synthesized. The probe, with an α,β‐unsaturated ketone conjugated into the coumarin fluorophore, was synthesized by 4 steps from the commercial material 4‐diethylamino salicylaldehyde. The electron push‐pull effect is enhanced with the N,N‐diethyl amino as the electron donor and the carbonyl as the electron acceptor. Both the absorption (463 nm) and emission (616 nm) maxima of CSP are in the visible wavelength region with a Stokes shift of about 150 nm, which ensures CSP a potential probe for biological application. Under near physiological conditions, significant fluorescence enhancement of CSP was observed upon the addition of some saccharides, namely, D‐sorbitol, D‐fructose, D‐glucose, D‐mannose and D‐galactose. The probe showed relatively high sensitivity towards D‐fructose and D‐sorbitol, and their detection limits were 0.05 mmol/L and 0.1 mmol/L, respectively.  相似文献   

5.
A novel chiral ionic liquid functionalized β‐cyclodextrin, 6‐O‐2‐hydroxpropyltrimethylammonium‐β‐cyclodextrin tetrafluoroborate ([HPTMA‐β‐CD][BF4]), was synthesized and used as a chiral selector in capillary electrophoresis. [HPTMA‐β‐CD][BF4] not only increased the solubility in aqueous buffer in comparison with the parent compound, but also provided a stable reversal electroosmotic flow, and the enantioseparation of eight chiral drugs was examined in phosphate buffer containing [HPTMA‐β‐CD][BF4] as the chiral selector. The effects of the [HPTMA‐β‐CD][BF4] concentration and the background electrolyte pH were studied. Moreover, the chiral separation abilities of β‐CD and [HPTMA‐β‐CD][BF4] were compared and possible mechanisms for the chiral recognition of [HPTMA‐β‐CD][BF4] are discussed. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

6.
The anthraquinone derivatives T‐x‐x ( x = 2, 4, and 8), possessing two cellobiosyl, cellotetraosyl, and cellooctaosyl chains, respectively, C‐glycosidically bonded at C(1) and C(8) were synthesised as potential mimics of cellulose I. The anthraquinone template enforces a parallel orientation of the cellodextrin chains at a distance corresponding to the one between the crystallographically independent chains of cellulose I, and the ethynyl and buta‐1,3‐diynyl linker units ensure an appropriate phase shift between them. The H‐bonding of the T‐x‐x mimics was analysed and compared to the one of the mono‐chained analogues T‐x and of the known cellulose II mimics N‐x‐x and N‐x where one or two cellodextrin chains are O‐glycosidically bonded to naphthalene‐1,8‐diethanol, or to naphthalene‐1‐ethanol. The OH signals of T‐x and T‐x‐x in solution in (D6)DMSO were assigned on the basis of DQFCOSY, HSQC, and TOCSY (only of T‐4, T‐4‐4 , and T‐8‐8 ) spectra and on a comparison with the spectra of N‐x and N‐x‐x. Hydrogen bonding was analysed on the basis of the chemical shift of OH groups and its temperature dependence, coupling constants, SIMPLE 1H‐NMR experiments, and ROESY spectra. T‐4‐4 and T‐8‐8 in (D6)DMSO appear to adopt a V‐shape arrangement of the cellosyl chains, avoiding inter‐chain H‐bond interactions. The well‐resolved solid‐state CP/MAS 13C‐NMR spectra of the mono‐chained T‐x ( x = 1, 2, 4, and 8) show that only T‐8 is a close mimic of cellulose II. While the solid‐state CP/MAS 13C‐NMR spectrum of the C1‐symmetric diglucoside T‐1‐1 is well‐resolved, the spectra of T‐2‐2 and T‐4‐4 show broad signals, and that of T‐8‐8 is rather well resolved. The spectrum of T‐8‐8 resembles that of cellulose Iβ. A comparison of the X‐ray powder‐diffraction spectra of T‐8‐8 and T‐8 with those of celluloses confirms that T‐8‐8 is a H‐bond mimic of cellulose I and T‐8 one of cellulose II. Surprisingly, there is little difference between the CP/MAS 13C‐NMR spectra of the acetyl protected mono‐chained C‐glycosylated anthraquinone derivatives A‐x and the double‐chained A‐x‐x ( x = 2, 4, and 8). The spectra of A‐4 and A‐4‐4 resemble strongly the one of cellulose triacetate I ( CTA I ). The (less well‐resolved) spectra of the cellooctaosides A‐8 and A‐8‐8 , however, resemble the one of CTA II . The similarity between the solid‐state CP/MAS 13C‐NMR spectra of A‐4 and A‐4‐4 to the one of CTA I , and of A‐8 and A‐8‐8 to the one of CTA II is opposite to the observations in the acetylated cellodextrin series. The mono‐chained A‐x cellulose triacetate mimics 21 ( A‐2 ), 32 ( A‐4 ), and 55 ( A‐8 ) were synthesised by Sonogashira coupling of the cellooligosyl‐ethynes 15, 28 , and 50 , followed by selective deacetylation. Complete deacetylation provided the corresponding T‐x mimics. The double‐chained A‐x‐x mimics 24 ( A‐2‐2 ), 35 ( A‐4‐4 ), and 58 ( A‐8‐8 ) were prepared from A‐x by triflation and Sonogashira coupling with the cellosyl‐buta‐1,3‐diynes 19, 31 , and 53 . Their deacetylation provided the corresponding T‐x‐x mimics 25, 36 , and 59 . The cellooligosyl‐ethynes and cellooligosyl‐buta‐1,3‐diynes required for the Sonogashira coupling were prepared by stepwise glycosylation of the partially O‐benzylated β‐cellobiosyl‐ethyne and β‐cellobiosyl‐buta‐1,3‐diyne 13 and 17 , respectively, with the cellobiosyl donor 2 and the cellohexaosyl donor 47 .  相似文献   

7.
β‐Thalassemia is one of the most common inherited disorders and is widely distributed throughout the world. Owing to severe deficiencies in red blood cell production, blood transfusion is required to correct anemia for normal growth and development but causes additional complications owing to iron overload. The aim of this study is to quantify the biometal dysregulations in β‐thalassemia patients as compared with healthy controls. A total of 17 elements were analyzed in serum samples of β‐thalassemia patients and healthy controls using ICP‐MS followed by chemometric analyses. Out of these analyzed elements, 14 showed a significant difference between healthy and disease groups at p < 0.05 and fold change >3. A PLS‐DA model revealed an excellent separation with 89.8% sensitivity and 97.2% specificity and the overall accuracy of the model was 92.2%. This metallomic study revealed that there is major difference in metallomic profiling of β‐thalassemia patients specifically in Co, Mn, Ni, V and Ba, whereas the fold changes in Co, Mn, V and Ba were found to be greater than that in Fe, providing evidence that, in addition to Fe, other metals are also altered significantly and therefore chelation therapy for other metals may also needed in β‐thalassemia patients.  相似文献   

8.
From the leaves of Solanum campaniforme, two new spirosolane alkaloids β‐acetoxyl‐(25S)‐22βN‐spirosol‐4‐en‐3‐one (1) and β‐hydroxyl‐(25S)‐22βN‐spirosol‐4‐en‐3‐one (4) were isolated along with two other known alkaloids of the same class (25S)‐22βN‐spirosol‐1,4‐dien‐3‐one (2) and (25S)‐22βN‐spirosol‐4‐en‐3‐one (3), which are reported for the first time as natural products. The structures of all alkaloids were established after an extensive analysis by 1D and 2D NMR spectroscopy (COSY, HSQC, HMBC and NOESY) as well as HRESIMS. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

9.
Five new triterpenoid saponins, including 3‐Oβ‐d ‐galactopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)]‐β‐d ‐glucuronopyranosyl quillaic acid 28‐Oβ‐d ‐glucopyranosyl‐(1→3)‐β‐d ‐xylopyranosyl‐(1→4)‐α‐l ‐rhamnopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)‐(4‐O‐acetyl)‐β‐d ‐quinovopyranosyl‐(1→4)]‐β‐d ‐fucopyranoside ( 1 ), 3‐Oβ‐d ‐galactopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)]‐β‐d ‐glucuronopyranosyl quillaic acid 28‐O‐(6‐O‐acetyl)‐β‐d ‐glucopyranosyl‐(1→3)‐[β‐d ‐xylopyranosyl‐(1→4)]‐α‐l ‐rhamnopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)‐(4‐O‐acetyl)‐β‐d ‐quinovopyranosyl‐(1→4)]‐β‐d ‐fucopyranoside ( 2 ), 3‐Oβ‐d ‐galactopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)]‐β‐d ‐glucuronopyranosyl quillaic acid 28‐Oβ‐d ‐xylopyranosyl‐(1→4)‐α‐l ‐rhamnopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)‐(4‐O‐acetyl)‐β‐d ‐quinovopyranosyl‐(1→4)]‐β‐d ‐fucopyranoside ( 3 ), 3‐Oβ‐d ‐galactopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)]‐β‐d ‐glucuronopyranosyl quillaic acid 28‐Oβ‐d ‐glucopyranosyl‐(1→3)‐β‐d ‐xylopyranosyl‐(1→4)‐α‐l ‐rhamnopyranosyl‐(1→2)‐[(4‐O‐acetyl)‐β‐d ‐quinovopyranosyl‐(1→4)]‐β‐d ‐fucopyranoside ( 4 ), 3‐Oβ‐d ‐galactopyranosyl‐(1→2)‐[β‐d ‐xylopyranosyl‐(1→3)]‐β‐d ‐glucuronopyranosyl quillaic acid 28‐O‐(6‐O‐acetyl)‐β‐d ‐glucopyranosyl‐(1→3)‐[β‐d ‐xylopyranosyl‐(1→4)]‐α‐l ‐rhamnopyranosyl‐(1→2)‐[(4‐O‐acetyl)‐β‐d ‐quinovopyranosyl‐(1→4)]‐β‐d ‐fucopyranoside ( 5 ) together with two known congeners, saponariosides A ( 6 ) and B ( 7 ) were isolated from the roots of Saponaria officinalis L. Their structures were elucidated by extensive spectroscopic methods, including 1D‐ (1H, 13C) and 2D‐NMR (DQF‐COSY, TOCSY, HSQC, and HMBC) experiments, HR‐ESI‐MS, and acid hydrolysis.  相似文献   

10.
Seven new acyl glycosides, benzyl 5‐O‐vanilloyl‐β‐d ‐apiofuranosyl‐(1→6)‐β‐d ‐glucopyranoside ( 1 ), 4‐hydroxy‐3‐methoxyphenyl 5‐O‐syringoyl‐β‐d ‐apiofuranosyl‐(1→6)‐β‐d ‐glucopyranoside ( 2 ), isopentyl 5‐O‐syringoyl‐β‐d ‐apiofuranosyl‐(1→6)‐β‐d ‐glucopyranoside ( 3 ), 3,4,5‐trimethoxyphenyl 5‐O‐sinapoyl‐β‐d ‐apiofuranosyl‐(1→6)‐β‐d ‐glucopyranoside ( 4 ), 6‐methoxy‐7‐[(6‐O‐sinapoyl‐β‐d ‐glucopyranosyl)oxy]coumarin ( 5 ), 6‐methoxy‐7‐[(2‐O‐sinapoyl‐β‐d ‐glucopyranosyl)oxy]coumarin ( 6 ), and isopentyl β‐d ‐apiofuranosyl‐(1→6)‐[5‐O‐syringoyl‐β‐d ‐apiofuranosyl‐(1→2)]‐β‐d ‐glucopyranoside ( 7 ), were isolated from Chinese folk herb Erycibe obtusifolia. Their structures were elucidated on the basis of extensive spectroscopic analysis, including UV, IR, MS, and 1D‐ and 2D‐NMR techniques. Further, these compounds were evaluated against HCT‐8 (human colon carcinoma), Bel‐7402 (human liver carcinoma), BGC‐823 (human stomach carcinoma), A549 (human lung carcinoma), and A2780 (human ovarian carcinoma) cell lines, however, none of them exhibited a significant bioactivity (IC50 > 10 μm ).  相似文献   

11.
Three new pregnane glycosides, cynatroside A ( 1 ), cynatroside B ( 2 ), and cynatroside C ( 3 ), isolated from the roots of Cynanchum atratum (Asclepiadaceae), were characterized as 7β‐{[Oα‐L ‐cymaropyranosyl‐(1→4)‐Oβ‐D ‐digitoxopyranosyl‐(1→4)‐β‐D ‐oleandropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐6α‐hydroxy‐4b‐ methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 1 ), 7β‐{[Oβ‐D ‐cymaropyranosyl‐(1→4)‐Oα‐L ‐diginopyranosyl‐(1→4)‐β‐D ‐cymaropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐2,6α‐dihydroxy‐4b‐methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 2 ), and 7β‐{[Oα‐L ‐cymaropyranosyl‐(1→4)‐Oβ‐D ‐digitoxopyranosyl‐(1→4)‐β‐L ‐cymaropyranosyl]oxy}‐3,4,4a,4b,5,6,7,8,10,10a‐decahydro‐2,6α‐dihydroxy‐4b‐methyl‐2‐(2‐methyl‐3‐furyl)phenanthren‐1(2H)‐one ( 3 ), respectively. In addition, ten known constituents were identified, i.e., cynascyroside D ( 4 ), glaucoside C ( 5 ), glaucoside D ( 6 ), atratoside A ( 7 ), 2,4‐dihydroxyacetophenone ( 8 ), 4‐hydroxyacetophenone ( 9 ), syringic acid ( 10 ), azelaic acid ( 11 ), suberic acid ( 12 ), and succinic acid ( 13 ). Among these compounds, 1 – 4 significantly inhibit acetylcholinesterase activity.  相似文献   

12.
Aloe‐emodin, a natural polyphenolic anthraquinone, has shown various beneficial bioactivities in vitro. The aim of this study was to investigate the pharmacokinetics and metabolism of aloe‐emodin. Aloe‐emodin was intravenously and orally administered to rats. The concentrations of aloe‐emodin and rhein, a metabolite of aloe‐emodin, were determined by HPLC method prior to and after hydrolysis with β‐glucuronidase and sulfatase/β‐glucuronidase. The results showed that the systemic exposures of aloe‐emodin and its metabolites were ranked as aloe‐emodin glucuronides (G) > rhein sulfates (S) > aloe‐emodin > rhein and rhein G when aloe‐emodin was given intravenously. In contrast, when aloe‐emodin was administered orally, the parent form of aloe‐emodin was not absorbed per se, and the systemic exposures of its metabolites were ranked as aloe‐emodin G > rhein G > rhein. In conclusion, the metabolites of aloe‐emodin are more important than the parent form for the bioactivities in vivo. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

13.
Two new compounds, (?)‐(6aR,11aR)‐4‐methoxy‐8,9‐(methylenedioxy)pterocarpan 3‐Oβ‐D ‐glucopyranoside ( 1 ) and 5‐hydroxy‐7‐methoxyisoflavone 4′‐Oβ‐D ‐xylopyranosyl‐(1→6)‐β‐D ‐glucopyranoside ( 2 ), were isolated, together with 30 known compounds from the stems and leaves of Sophora flavescens Aition . Their structures were elucidated by extensive spectroscopic analysis, including HR‐ESI‐MS data. A preliminary comparison of phenolic metabolite profiles, based on the qualitative HPLC analysis, indicated that the composition of the roots and the aerial parts were significantly different.  相似文献   

14.
Heat shock proteins (HSPs) hsp70/hsc70, hsp90 and hsp96 were separated from mammalian cells and tissues on a gel obtained by the reaction of β‐mercaptoethanol with divinyl sulfone‐activated Sepharose CL‐6B (thiophilic gel or T‐gel). Hsp90 revealed a much higher affinity towards the T‐gel than the other HSPs. One‐step thiophilic interaction chromatography of proteins resulted in a more than 80% purity and 85% yield of hsp90. Based on this observation, a simple and efficient method for the purification of hsp90 and a procedure for the simultaneous purification of several HSPs (hsp70/hsc70, hsp90 and hsp96) using thiophilic interaction chromatography was developed. All the HSPs were recovered with a high yield and purity (90–99%). The results indicated that the thiophilic gel is a highly efficient affinity matrix for the purification of hsp90 and can be used in the protocols of purification of different HSPs from cells and tissues of various animal species. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

15.
Two new epimeric chlorinated withaphysalins, rel‐(4β,5β,6α,18S,22R)‐ and rel‐(4β,5β,6α,18R,22R)‐6‐chloro‐18,20‐epoxy‐18‐ethoxy‐4,5‐dihydroxy‐1‐oxowitha‐2,24‐diene‐26,22‐lactone ( 1 and 2 resp.), together with the new rel‐(4β,5β,6α,18R,22R)‐6‐chloro‐18,20‐epoxy‐4,5‐dihydroxy‐18‐methoxy‐1‐oxowitha‐2,24‐diene‐26,22‐lactone ( 3 ) and rel‐(3β,4β,5β,6β,18R,22R)‐5,6:18,20‐diepoxy‐3,18‐diethoxy‐4‐hydroxy‐1‐oxowith‐24‐ene‐26,22‐lactone ( 4 ) were isolated from the leaves of Acnistus arborescens and named withaphysalins T–W, respectively. The final structures and the complete 1H‐ and 13C‐NMR assignments of the three chlorowithaphysalins 1 – 3 were performed by means of HR‐ESI‐MS and 1D‐ and 2D‐NMR experiments, including COSY, HSQC, and HMBC, beside comparison with spectral data of analogous compounds from the literature. The structure of 4 was also confirmed by means of a single‐crystal X‐ray diffraction analysis.  相似文献   

16.
Current therapeutic interventions in bone defects are mainly focused on finding the best bioactive materials for inducing bone regeneration via activating the related intracellular signaling pathways. Integrins are trans‐membrane receptors that facilitate cell‐extracellular matrix (ECM) interactions and activate signal transduction. To develop a suitable platform for supporting human bone marrow mesenchymal stem cells (hBM‐MSCs) differentiation into bone tissue, electrospun poly L‐lactide (PLLA) nanofiber scaffolds were coated with nano‐hydroxyapatite (PLLA/nHa group), gelatin nanoparticles (PLLA/Gel group), and nHa/Gel nanoparticles (PLLA/nHa/Gel group) and their impacts on cell proliferation, expression of osteoblastic biomarkers, and bone differentiation were examined and compared. MTT data showed that proliferation of hBM‐MSCs on PLLA/nHa/Gel scaffolds was significantly higher than other groups (P < .05). Alkaline phosphatase activity was also more increased in hBM‐MSCs cultured under osteogenic media on PLLA/nHa/Gel scaffolds compared to others. Gene expression evaluation confirmed up‐regulation of integrin α2β1 as well as the osteogenic genes BGLAP, COL1A1, and RUNX2. Following use of integrin α2β1 blocker antibody, the protein level of integrin α2β1 in cells seeded on PLLA/nHa/Gel scaffolds was decreased compared to control, which confirmed that most of the integrin receptors were bound to gelatin molecules on scaffolds and could activate the integrin α2β1/ERK axis. Collectively, PLLA/nHa/Gel scaffold is a suitable platform for hBM‐MSCs adhesion, proliferation, and osteogenic differentiation in less time via activating integrin α2β1/ERK axis, and thus it might be applicable in bone tissue engineering.  相似文献   

17.
Bacteriophage (phage) proteins have been analyzed previously with matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF MS). However, analysis of phage major capsid proteins (MCPs) has been limited by the ability to reproducibly generate ions from MCP monomers. While the acidic conditions of MALDI‐TOF MS sample preparation have been shown to aid in disassembly of some phage capsids, many require further treatment to successfully liberate MCP monomers. The findings presented here suggest that β‐mercaptoethanol reduction of the disulfide bonds linking phage MCPs prior to mass spectrometric analysis results in significantly increased MALDI‐TOF MS sensitivity and reproducibility of Yersinia pestis‐specific phage protein profiles. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

18.
Diets rich in high fructose and salt are increasingly popular in our daily life. A combination consumption of excessive fructose and salt can induce insulin resistance (IR) and hypertension (HT), which are major public health problems around the world. However, the effects of high fructose and salt on systematic metabonome remain unknown, which is very important for revealing the molecular mechanism of IR and HT induced by this dietary pattern. The metabolic profiling in urine, plasma, and fecal extracts from high fructose and salt‐fed rats was investigated by use of 1H nuclear magnetic resonance (NMR)‐based metabonomics approach in this study. Multivariate analysis of NMR data showed the effects of high fructose and salt on the global metabonome. The metabolite analysis in urine and fecal extracts showed the time‐dependent metabolic changes, which displayed metabonomic progression axes from normal to IR and HT status. The changes of 2‐oxoglutarate, creatine and creatinine, citrate, hippurate, trimethylamine N‐oxide (TMAO), and betaine in urine, together with gut microbiota disorder in feces, were observed at the preliminary formation stage of IR and HT (fourth week). At the severe stage (eighth week), the previously mentioned metabolic changes were aggravated, and the changes of lipid and choline metabolism in plasma suggested the increased risk of cardiovascular diseases. These findings provide an overview of biochemistry consequences of high fructose and salt feeding and comprehensive insights into the progression of systematic metabonome for IR and HT induced by this dietary pattern. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

19.
Guided by in vitro immunological tests, three immunomodulating steroidal glycosides, stemucronatosides A ( 1 ), B ( 2 ), and C ( 3 ), were isolated from the roots of Stephanotis mucronata. On the basis of chemical evidence and extensive spectroscopic methods including 1D and 2D NMR, their structures were determined as 12‐O‐deacetylmetaplexigenin 3‐[O‐6‐deoxy‐3‐O‐methyl‐β‐D ‐allopyranosyl‐(1→4)‐Oβ‐D ‐cymaropyranosyl‐(1→4)‐β‐D ‐cymaropyranoside], 12‐O‐deacetylmetaplexigenin 3‐[Oβ‐D ‐thevetopyranosyl‐(1→4)‐Oβ‐D ‐cymaropyranosyl‐(1→4)‐β‐D ‐cymaropyranoside], and metaplexigenin 3‐[Oβ‐D ‐glucopyranosyl‐(1→4)‐O‐6‐deoxy‐3‐O‐methyl‐β‐D ‐allopyranosyl‐(1→4)‐Oβ‐D ‐cymaropyranosyl‐(1→4)‐β‐D ‐cymaropyranoside], respectively. These compounds showed immunomodulating activities in vitro.  相似文献   

20.
Three new medicagenic acid saponins, micranthosides A–C ( 1 – 3 ), were isolated from the roots of Polygala micrantha Guill . & Perr ., along with six known presenegenin saponins. Their structures were elucidated on the basis of extensive 1D‐ and 2D‐NMR experiments (1H, 13C, DEPT, COSY, TOCSY, NOESY, HSQC, and HMBC) and mass spectrometry as 3‐Oβ‐D ‐glucopyranosylmedicagenic acid 28‐[Oβ‐D ‐galactopyranosyl‐(1→4)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐D ‐fucopyranosyl] ester ( 1 ), 3‐Oβ‐D ‐glucopyranosylmedicagenic acid 28‐[O‐6‐O‐acetyl‐β‐D ‐galactopyranosyl‐(1→4)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐D ‐fucopyranosyl] ester ( 2 ), and 3‐O‐{Oβ‐D ‐glucopyranosyl‐(1→3)‐O‐[β‐D ‐glucopyranosyl‐(1→6)]‐β‐D ‐glucopyranosyl}medicagenic acid 28‐{Oβ‐D ‐apiofuranosyl‐(1→3)‐Oβ‐D ‐xylopyranosyl‐(1→4)‐O‐[β‐D ‐apiofuranosyl‐(1→3)]‐Oα‐L ‐rhamnopyranosyl‐(1→2)‐β‐D ‐fucopyranosyl} ester ( 3 ). Compounds 1 – 3 were evaluated against HCT 116 and HT‐29 human colon cancer cells, but they did not show any cytotoxicity.  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