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.
[Pt(COD)Cl2] (COD=1,5-cyclooctadiene) is a versatile starting material for the synthesis of Pt(II) compounds. The preparations of the new compounds [Pt(COD)Cl(NO3)], [Pt(COD)(NO3)2] and [Pt(PPh3)3(NO3)](NO3) and also of the known compounds cis[Pt(PPh3)2Cl2], cis [Pt(PPh3)2Cl(NO3)], cis[Pt(PPh3)2(NO3)2] and [Pt(PPh3)3Cl](NO3)are reported. The compounds are characterized by elemental analysis, 31P{1H} NMR spectroscopy and IR spectroscopy.  相似文献   

2.
《Life sciences》1986,39(16):1457-1464
Incubation of racemic methylenedioxyamphetamine (MDA) or methylenedioxymethamphetamine (MDMA) with rat hepatic microsomes, in the presence of NADPH, generated a spectrally observed inhibitory complex with cytochrome P-450. The complex inhibited product formation from MDA and MDMA as well as other P-450 dependent reactions such as benzphetamine demethylation and CO binding. In the absence of NADPH, MDMA and MDA generated type I and type IIa difference spectra, respectively, suggesting differences in their binding to the enzyme active site. The N-demethylation of MDMA was partially inhibited by methimazole suggesting involvement of the hepatic flavin-containing monooxygenase.  相似文献   

3.
Detailed electrostatic potential (ESP) analyses were performed to compare the directionality of halogen bonds with those of hydrogen bonds and lithium bonds. To do this, the interactions of HOOOH with the molecules XF (X?=?Cl, Br, H, Li) were investigated. For each molecule, the percentage of the van der Waals (vdW) molecular surface that intersected with the ESP surface was used to roughly quantify the directionality of the halogen/hydrogen/lithium bond associated with the molecule. The size of the region of intersection was found to increase in the following order: ClF?<?BrF?<?HF?<?LiF. The maximum ESP in the region of intersection, V S, max, was observed to become more positive according to the sequence ClF?<?BrF?<?HF?<?LiF. For ClF and BrF, the positive electrostatic potential was concentrated in a very small region of the vdW molecular surface. On the other hand, for HF and LiF, the positive electrostatic potential was more diffusely scattered across the vdW surface than for ClF and BrF. Also, the optimized geometries of the dipolymers HOOOH···?XF (X?=?Cl, Br, H, Li) indicated that halogen bonds are more directional than hydrogen bonds and lithium bonds, consistent with the results of ESP analyses.
Graphical abstract Electrostatic potential (ESP) contour maps in the xz plane of ClF and BrF
  相似文献   

4.
This article describes the synthesis of (3 ′S) and (3 ′R)-3 ′-amino-3 ′-deoxy pyranonucleosides and their precursors (3 ′S) and (3 ′R)-3 ′-azido-3 ′-deoxy pyranonucleosides. Azidation of 1,2:5,6-di-O-isopropylidene-3-O-toluenesulfonyl-α-D-allofuranose followed by hydrolysis and subsequent acetylation afforded 3-azido-3-deoxy-1,2,4,6-tetra-O-acetyl-D-glucopyranose, which upon coupling with the proper silylated bases, deacetylation, and catalytic hydrogenation, obtained the target 3 ′-amino-3 ′-deoxy-β-D-glucopyranonucleosides. The desired 1-(3 ′-amino-3 ′-deoxy-β-D-allopyranosyl)5-fluorouracil was readily prepared from the suitable imidazylate sugar after azidation followed by a protection/deprotection sequence and reduction of the unprotected azido precursor. No antiviral activity was observed for the novel nucleosides. Moderate cytostatic activity was recorded for the 5-fluorouracil derivatives.  相似文献   

5.
The RGD (Arg-Gly-Asp) binding integrins α(v)β(3) and α(IIb)β(3) are integral components of various pathological and physiological processes, including tumor angiogenesis, osteoclast function, and thrombus formation. Because of this, there is interest in identifying novel compounds and proteins binding to these receptors as well as investigating the mechanism of these interactions. In this article, we describe the development and validation of competition binding assays for determining the affinity of test compounds to α(v)β(3) and α(IIb)β(3) integrin. Assays were successfully developed for each receptor, and the affinity of known compounds was comparable to published results. However, the inability of binding between α(IIb)β(3) integrin and the labeled echistatin protein ligand to reach equilibrium resulted in an assay that did not meet the assumptions of the competition binding model. Nevertheless, there was good agreement between this assay and known literature values, and intra- and interassay variability was acceptable. Binding by conformation-specific antibodies provided evidence that solid-phase bound α(IIb)β(3) receptor was in an activated conformation. This study also demonstrated that current models and methods for determining receptor affinity are simplistic and fail to account for common receptor-ligand interactions such as nondissociable interactions and varying receptor activation states.  相似文献   

6.
Two 3-(7′-theophyllyl)glycals, (IV) and (V), were synthesized by fusion of theophylline and the appropriate glycals in the presence of p-toluenesulfonic acid. The structure and stereochemistry of the glycals were determined mainly from NMR analysis of their dihydro and 1,6-anhydro derivatives.  相似文献   

7.
A series of novel 2,5-bis(3′-indolyl)furans and 3,5-bis(3′-indolyl)isoxazoles were synthesized as antitumor agents. The antiproliferative activity was evaluated in vitro toward diverse human tumor cell lines. Initially 5 isoxazoles and 3 furan derivatives were tested against a panel of 10 human tumor cell lines and the most active derivatives 3c and 4a were selected to be evaluated in an extended panel of 29 cell lines. By exhibiting mean IC50 values of 17.4 μg/mL (3a) and 20.5 μg/mL (4c), in particular 4c showed a high level of tumor selectivity toward the 29 cell lines.  相似文献   

8.
Tea catechins, (–)-epigallocatechin-3-gallate (EGCg) and (–)-epigallocatechin (EGC), have been reported to suppress oxidation of plasma low density lipoprotein (LDL) in vitro. If dietary catechins can be efficiently incorporated into human blood plasma, anti-atherosclerotic effects in preventing oxidative modification of LDL would be expected. In this study, a newly developed chemiluminescence detection-high pressure liquid chromatography (CL-HPLC) method for measuring plasma catechins was used and the incorporation of EGCg and EGC into human plasma was investigated. Healthy subjects orally ingested 3, 5, or 7 capsules of green tea extract (corresponding to 225, 375, and 525 mg EGCg and 7.5, 12.5, and 17.5 mg EGC, respectively). The plasma EGCg and EGC concentrations before the administration were all below the detection limit (< 2 pmol/ml), but 90 min after, significantly and dose-dependently increased to 657, 4300, and 4410 pmol EGCg/ml, and 35, 144, and 255 pmol EGC/ml, in the subjects who received 3, 5, and 7 capsules, respectively. Both EGCg and EGC levels detected in plasma corresponded to 0.2–2.0% of the ingested amount. Catechin intake had no effect on the basal level of endogenous antioxidants (α-tocopherol, β-carotene, and lycopene) or of lipids in plasma. These results suggested that drinking green tea daily would contribute to maintain plasma catechin levels sufficient to exert antioxidant activity against oxidative modification of lipoproteins in blood circulation systems.  相似文献   

9.
Early studies led to the identification of 3β-(4-methoxyphenyl)tropane-2β-carboxylic acid methyl ester (5) with high affinity at the DAT (IC50 = 6.5 nM) and 5-HTT (Ki = 4.3 nM), while having much less affinity at the NET (Ki = 1110 nM). In the present study, we replaced the 4′-methoxy group of the 3β-phenyl ring with a bioisosteric 4′-methylthio group to give 7a. We also synthesized a number of 3β-(4-alkylthiophenyl)tropanes 7be, 3β-(4-methylsulfinylphenyl) and 3β-(4-methylsulfonylphenyl)tropane analogues 7fh as well as the 3β-(4-alkylthiophenyl)nortropane derivatives 811 to further characterize the structure–activity relationship of this type of compound for binding at monoamine transporters. With exception of the 4′-methylsulfonyl analogue 7h, all the tested compounds possessed high binding affinities at the 5-HTT. The Ki values ranged from 0.19 nM to 49 nM. The 3β-(4-methylthiophenyl)tropane 7a and its N-(3-fluoropropyl) analogue 9a and N-allyl analogue 10a are the most selective compounds for the 5-HTT over the NET (NET/5-HTT = 314–364) in the series. However, none of the compounds showed selectivity similar to 5 for both the DAT and 5-HTT relative to the NET. This study provided useful SAR information for rational design of potent and selective monoamine transporter inhibitors.  相似文献   

10.
三冠大孢属Tricristatispora是我国西部地区上三叠统特征的常见化石。此前分别命名为Calamospora tricristata Li,1974,Tricristatispora aphela Liu,1981和Viburamegaspora orientalis Zhang,1984的几个种实属具相同特征的同种化石,后2个种名为前1种的晚出同义名,应予废弃。这里将种名tricrist  相似文献   

11.
12.
方宗杰 《古生物学报》2022,61(4):662-663
1993年, 于菁珊和董国义为辽宁上三叠统的非海相双壳类化石建立了一个新属辽宁蚌Liaoningia Yu and Dong, 1993, 以Liaoningia opima Yu and Dong, 1993作为模式种。然而, 早在1979年, 邢裕盛和刘桂芝就已为辽宁晚前寒武纪南关岭组所产的化石标本创建了新属名辽宁水母属Liaoningia Xing and Liu, 1979, 模式种是Liaoningia fuxianensis Xing and Liu, 1979。然而, 多数学者反对将它视为水母类化石, 有人怀疑辽宁水母属是假化石, 迄今学术界并无统一意见, 故本文暂将它归为疑问化石(Problematica)。根据“国际动物命名法规”, 我们提出一个新的属名Liaoningoconcha nom. nov., 用以替代Liaoningia Xing and Liu, 1979的次同义名Liaoningia Yu and Dong, 1993, 仍使用于菁珊和董国义在1993年指定的模式种, 中文译名不变。  相似文献   

13.
The racemic total synthesis of elegansidiol, farnesiferol B, and farnesiferol D has been obtained following a Diels–Alder approach. Gillman addition, cross metathesis reaction are the other key steps involved in the target synthesis.  相似文献   

14.
The synthesis of new zirconium and hafnium mixed ligand phthalocyanine complexes PcM(β-ketoester)2, where M-Zr (IV), Hf (IV); Pc - the dianion of phthalocyanine, and β-ketoester - the out planed ligand, is reported. The obtained complexes are characterized by 1H NMR, IR, UV-Vis spectroscopy and cyclic voltammetry. 1H NMR and elemental analysis confirm the substitution of two Cl atoms for two β-ketoester fragments to the central atom of the macrocycle. The data of 1H NMR, UV-Vis spectroscopy have allowed us to conclude that two β-ketoester ligands are in the cis geometry to the phthalocyanine plane. X-ray crystallography for bis(isopropyl 3-oxobutanoato)hafnium(IV)phthalocyanine confirms this conclusion. The central macrocycle of the phthalocyanine ligand is not exactly planar (deviations from the least-square plane exceed 0.15 Å) and has the conformation of an essentially flattened crown. The Hf(1) atom is 1.349(3) above this least-square plane. Cyclic voltammetry investigation shows that the introduction of two β-ketoester ligands to the central atom of phthalocyanine complex leads to both chemical and electrochemical stabilization of the whole Pc system.  相似文献   

15.
根据一件产自云南罗平中三叠世关岭组Ⅱ段的新标本并结合产自相同地点和地层中的模式标本对纤细滇美龙(Dianmeisaurus gracilis Shang & Li,2015)进行了详细研究.原模式标本暴露其腹而,而新标本暴露其背面,两者互相补充提供了更完整、精确的纤细滇美龙解剖学信息.新材料揭示该种具有非常短小的吻部,眶前区的长度不仅短于眶后区长度,甚至短于眼眶的长度;外鼻孔小且位置靠前,即鼻孔前区的长度短于鼻孔后缘与眼眶前缘之间的距离;由两额骨构成的眼眶间隔非常狭窄,宽度小于顶骨平台宽度的1/3;额骨前后两端均具渐尖的突起;顶骨后部不收缩,顶骨平台后缘呈深V型.补充了新信息和包含更多属种(如Dawazisaurus)后的系统发育学分析支持了之前滇美龙和滇东龙互为姊妹群的结论,同时它们和马家山龙、滇肿龙、贵州龙和大洼子龙一起构成了一个仅由中国的属种组成的单系类群.与欧洲肿肋龙类群(Dactylosaurus,Anarosaurus,Serpianosaurus和Neusticosaurus)相比,这一单系类群与幻龙类有更近的亲缘关系.  相似文献   

16.
Glycoside hydrolase family 1 (GH1) β-glucosidases play roles in many processes in plants, such as chemical defense, alkaloid metabolism, hydrolysis of cell wall-derived oligosaccharides, phytohormone regulation, and lignification. However, the functions of most of the 34 GH1 gene products in rice (Oryza sativa) are unknown. Os3BGlu6, a rice β-glucosidase representing a previously uncharacterized phylogenetic cluster of GH1, was produced in recombinant Escherichia coli. Os3BGlu6 hydrolyzed p-nitrophenyl (pNP)-β-d-fucoside (kcat/Km = 67 mm−1 s−1), pNP-β-d-glucoside (kcat/Km = 6.2 mm−1 s−1), and pNP-β-d-galactoside (kcat/Km = 1.6 mm−1s−1) efficiently but had little activity toward other pNP glycosides. It also had high activity toward n-octyl-β-d-glucoside and β-(1→3)- and β-(1→2)-linked disaccharides and was able to hydrolyze apigenin β-glucoside and several other natural glycosides. Crystal structures of Os3BGlu6 and its complexes with a covalent intermediate, 2-deoxy-2-fluoroglucoside, and a nonhydrolyzable substrate analog, n-octyl-β-d-thioglucopyranoside, were solved at 1.83, 1.81, and 1.80 Å resolution, respectively. The position of the covalently trapped 2-F-glucosyl residue in the enzyme was similar to that in a 2-F-glucosyl intermediate complex of Os3BGlu7 (rice BGlu1). The side chain of methionine-251 in the mouth of the active site appeared to block the binding of extended β-(1→4)-linked oligosaccharides and interact with the hydrophobic aglycone of n-octyl-β-d-thioglucopyranoside. This correlates with the preference of Os3BGlu6 for short oligosaccharides and hydrophobic glycosides.β-Glucosidases (EC 3.2.1.21) have a wide range of functions in plants, including acting in cell wall remodeling, lignification, chemical defense, plant-microbe interactions, phytohormone activation, activation of metabolic intermediates, and release of volatiles from their glycosides (Esen, 1993). They fulfill these roles by hydrolyzing the glycosidic bond at the nonreducing terminal glucosyl residue of a glycoside or an oligosaccharide, thereby releasing Glc and an aglycone or a shortened carbohydrate. The aglycone released from the glycoside may be a monolignol, a toxic compound, or a compound that further reacts to release a toxic component, an active phytohormone, a reactive metabolic intermediate, or a volatile scent compound (Brzobohatý et al., 1993; Dharmawardhama et al., 1995; Reuveni et al., 1999; Lee et al., 2006; Barleben et al., 2007; Morant et al., 2008). Indeed, the wide range of glucosides of undocumented functions found in plants suggests that many β-glucosidase functions may remain to be discovered.Plant β-glucosidases fall into related families that have been classified as glycosyl hydrolase (GH) families GH1, GH3, and GH5 (Henrissat, 1991; Coutinho and Henrissat, 1998, 1999). Of these, GH1 has been most thoroughly documented and shown to comprise a gene family encoding 40 putative functional GHs in Arabidopsis (Arabidopsis thaliana) and 34 in rice (Oryza sativa) in addition to a few pseudogenes (Xu et al., 2004; Opassiri et al., 2006). In addition to β-glucosidases, plant GH1 members include β-mannosidases (Mo and Bewley, 2002), β-thioglucosidases (Burmeister et al., 1997), and disaccharidases such as primeverosidase (Mizutani et al., 2002) as well as hydroxyisourate hydrolase, which hydrolyzes the internal bond in a purine ring rather than a glycosidic bond (Raychaudhuri and Tipton, 2002). The specificity for the glycone in GH1 enzymes varies. Some enzymes are quite specific for β-d-glucosides or β-d-mannosides, while many accept either β-d-glucosides or β-d-fucosides, and some also hydrolyze β-d-galactosides, β-d-xylosides, and α-l-arabinoside (Esen, 1993). However, most GH1 enzymes are thought to hydrolyze glucosides in the plant, and it is the aglycone specificity that determines the functions of most GH1 enzymes.Aglycone specificity of GH1 β-glucosidases ranges from rather broad to absolutely specific for one substrate and is not obvious from sequence similarity. For instance, maize (Zea mays) ZmGlu1 β-glucosidase hydrolyzes a range of glycosides, including its natural substrate, 2-O-β-d-glucopyranosyl-4-dihydroxy-1,4-benzoxazin-3-one (DIMBOAGlc), but not dhurrin, whereas sorghum (Sorghum bicolor) Dhr1, which is 72% identical to ZmGlu1, only hydrolyzes its natural cyanogenic substrate dhurrin (Verdoucq et al., 2003). Similarly, despite sharing over 80% amino acid sequence identity, the legume isoflavonoid β-glucosidases dalcochinase from Dalbergia cochinchinensis and Dnbglu2 from Dalbergia nigrescens hydrolyze each other''s natural substrate very poorly (Chuankhayan et al., 2007). Thus, small differences in the amino acid sequence surrounding the active site may be expected to account for significant differences in substrate specificity.GH1 is classified in GH clan A, which consists of GH families whose members have a (β/α)8-barrel structure with the catalytic acid/base on strand 4 of the β-barrel and the catalytic nucleophile on strand 7 (Henrissat et al., 1995; Jenkins et al., 1995). As such, all GH1 enzymes have similar overall structures, but it has been noted that four variable loops at the C-terminal end of the β-barrel strands, designated A, B, C, and D, account for much of the differences in the active site architecture (Sanz-Aparicio et al., 1998). The similar structures with great diversity in substrate specificity make plant GH1 enzymes an ideal model system to investigate the structural basis of substrate specificity. To date, seven plant β-glucosidase structures have been reported, including three closely related chloroplastic enzymes from maize (Czjzek et al., 2000, 2001), sorghum (Verdoucq et al., 2004), and wheat (Triticum aestivum; Sue et al., 2006), the cytoplasmic strictosidine β-glucosidase from Rauvolfia serpentine (Barleben et al., 2007), and the secreted enzymes white clover (Trifolium repens) cyanogenic β-glucosidase (Barrett et al., 1995), white mustard (Sinapsis alba) myrosinase (thioglucosidase; Burmeister et al., 1997), and rice Os3BGlu7 (BGlu1; Chuenchor et al., 2008). These enzymes hydrolyze substrates with a range of structures, but they cannot account for the full range of β-glucosidase substrates available in plants, and determining the structural differences that bring about substrate specificity differences in even closely related GH1 enzymes has proven tricky (Verdoucq et al., 2003, 2004; Sue et al., 2006; Chuenchor et al., 2008).Amino acid sequence-based phylogenetic analysis of GH1 enzymes encoded by the rice genome showed that there are eight clusters containing both rice and Arabidopsis proteins that are more closely related to each other than they are to enzymes from the same plants outside the clusters (Fig. 1; Opassiri et al., 2006). In addition, there are a cluster of sixteen putative β-glucosidases and a cluster of myrosinases in Arabidopsis without any closely related rice counterparts. Comparison with characterized GH1 enzymes from other plants reveals other clusters of related enzymes not found in rice or Arabidopsis, including the chloroplastic enzymes, from which the maize, sorghum, and wheat structures are derived, and the cytoplasmic metabolic enzymes, from with the strictosidine hydrolase structure is derived (Fig. 1). Therefore, although the known structures provide good tools for molecular modeling of plant enzymes, most rice and Arabidopsis GH1 enzymes lack a close correspondence in sequence and functional evolution to these structures, suggesting that the variable loops that determine the active site may be different. It would be useful, therefore, to know the structures and substrate specificities of representative members of each of the eight clusters seen in rice and Arabidopsis. To begin to acquire this information, we have expressed Os3BGlu6, a member of cluster At/Os 1 in Figure 1, characterized its substrate specificity, and determined its structure alone and in complex with a glycosyl intermediate and a nonhydrolyzable substrate analog.Open in a separate windowFigure 1.Simplified phylogenetic tree of the amino acid sequences of eukaryotic GH1 proteins with known structures and those of rice and Arabidopsis GH1 gene products. The protein sequences of the eukaryotic proteins with known structures are marked with four-character PDB codes for one of their structures, including Trifolium repens cyanogenic β-glucosidase (1CBG; Barrett et al., 1995), Sinapsis alba myrosinase (1MYR; Burmeister et al., 1997), Zea mays ZmGlu1 β-glucosidase (1E1F; Czjzek et al., 2000), Sorghum bicolor Dhr1 dhurrinase (1V02; Verdoucq et al., 2004), Triticum aestivum β-glucosidase (2DGA; Sue et al., 2006), Rauvolfia serpentina strictosidine β-glucosidase (2JF6; Barleben et al., 2007), and Oryza sativa Os3BGlu7 (BGlu1) β-glucosidase (2RGL; Chuenchor et al., 2008) from plants, along with Brevicoryne brassicae myrosinase (1WCG; Husebye et al., 2005), Homo sapiens cytoplasmic (Klotho) β-glucosidase (2E9M; Hayashi et al., 2007), and Phanerochaete chrysosporium (2E3Z; Nijikken et al., 2007), while those encoded in the Arabidopsis and rice genomes are labeled with the systematic names given by Xu et al. (2004) and Opassiri et al. (2006), respectively. One or two example proteins from each plant are given for each of the eight clusters of genes shared by Arabidopsis (At) and rice (Os) and the Arabidopsis-specific clusters At I (β-glucosidases) and At II (myrosinases), with the number of Arabidopsis or rice enzymes in each cluster given in parentheses. These sequences were aligned with all of the Arabidopsis and rice sequences in ClustalX (Thompson et al., 1997), the alignment was manually edited, all but representative sequences were removed, and the tree was calculated by the neighbor-joining method, bootstrapped with 1,000 trials, and then drawn with TreeView (Page, 1996). The grass plastid β-glucosidases, which are not represented in Arabidopsis and rice, are shown in the group marked “Plastid.” Percentage bootstrap reproducibility values are shown on internal branches where they are greater than 60%. Except those marked by asterisks, all external branches represent groups with 100% bootstrap reproducibility. To avoid excess complexity, those groups of sequences marked with asterisks are not monophyletic and represent more branches within the designated cluster than are shown. For a complete phylogenetic analysis of Arabidopsis and rice GH1 proteins, see Opassiri et al. (2006).  相似文献   

17.
18.
19.
Xue X  Fry SC 《Annals of botany》2012,109(5):873-886

Background and Aims

Horsetails (Equisetopsida) diverged from other extant eusporangiate monilophytes in the Upper Palaeozoic. They are the only monilophytes known to contain the hemicellulose mixed-linkage (1 → 3, 1 → 4)-β-d-glucan (MLG), whereas all land plants possess xyloglucan. It has been reported that changes in cell-wall chemistry often accompanied major evolutionary steps. We explored changes in hemicelluloses occurring during Equisetum evolution.

Methods

Hemicellulose from numerous monilophytes was treated with lichenase and xyloglucan endoglucanase. Lichenase digests MLG to di-, tri- and tetrasaccharide repeat-units, resolvable by thin-layer chromatography.

Key Results

Among monilophytes, MLG was confined to horsetails. Our analyses support a basal trichotomy of extant horsetails: MLG was more abundant in subgenus Equisetum than in subgenus Hippochaete, and uniquely the sister group E. bogotense yielded almost solely the tetrasaccharide repeat-unit (G4G4G3G). Other species also gave the disaccharide, whereas the trisaccharide was consistently very scarce. Tetrasaccharide : disaccharide ratios varied interspecifically, but with no consistent difference between subgenera. Xyloglucan was scarce in Psilotum and subgenus Equisetum, but abundant in subgenus Hippochaete and in the eusporangiate ferns Marattia and Angiopteris; leptosporangiate ferns varied widely. All monilophytes shared a core pattern of xyloglucan repeat-units, major XEG products co-chromatographing on thin-layer chromatography with non-fucosylated hepta-, octa- and nonasaccharides and fucose-containing nona- and decasaccharides.

Conclusions

G4G4G3G is the ancestral repeat-unit of horsetail MLG. Horsetail evolution was accompanied by quantitative and qualitative modification of MLG; variation within subgenus Hippochaete suggests that the structure and biosynthesis of MLG is evolutionarily plastic. Xyloglucan quantity correlates negatively with abundance of other hemicelluloses; but qualitatively, all monilophyte xyloglucans conform to a core pattern of repeat-unit sizes.  相似文献   

20.
Previous studies have shown that breast tissues and breast cell lines convert progesterone (P) to 5α-dihydroprogesterone (5αP) and 3α-dihydroprogesterone (3αHP) and that 3αHP suppresses, whereas 5αP promotes, cell proliferation and detachment. The objectives of the current studies were to determine if the 5αP- and 3αHP-induced changes in cell numbers are due to altered rates of mitosis and/or apoptosis, and if 3αHP and 5αP act on tumorigenic and non-tumorigenic cells, regardless of estrogen (E) and P receptor status. The studies were conducted on tumorigenic (MCF-7, MDA-MB-231, T47D) and non-tumorigenic (MCF-10A) human breast cell lines, employing several methods to assess the effects of the hormones on cell proliferation, mitosis, apoptosis and expression of Bcl-2, Bax and p21. In all four cell lines, 5αP increased, whereas 3αHP decreased cell numbers, [3H]thymidine uptake and mitotic index. Apoptosis was stimulated by 3αHP and suppressed by 5αP. 5αP resulted in increases in Bcl-2/Bax ratio, indicating decreased apoptosis; 3αHP resulted in decreases in Bcl-2/Bax ratio, indicating increased apoptosis. The effects of either 3αHP or 5αP on cell numbers, [3H]thymidine uptake, mitosis, apoptosis, and Bcl-2/Bax ratio, were abrogated when cells were treated simultaneously with both hormones. The expression of p21 was increased by 3αHP, and was unaffected by 5αP. The results provide the first evidence that 5αP stimulates mitosis and suppresses apoptosis, whereas 3αHP inhibits mitosis and stimulates apoptosis. The opposing effects of 5αP and 3αHP were observed in all four breast cell lines examined and the data suggest that all breast cancers (estrogen-responsive and unresponsive) might be suppressed by blocking 5αP formation and/or increasing 3αHP. The findings further support the hypothesis that progesterone metabolites are key regulatory hormones and that changes in their relative concentrations in the breast microenvironment determine whether breast tissues remain normal or become cancerous.  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