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 625 毫秒
1.
Platinum(II) complexes of the type [Pt(L)(cat)] ( 1 and 2 ), in which H2cat is catechol and L represents two 2‐(2‐pyridyl)benzimidazole ligands with 4,4‐difluoro‐4‐bora‐3a,4a‐diaza‐s‐indacene (BODIPY) pendants, were synthesized to achieve mitochondria‐targeted photocytotoxicity. The complexes showed strong absorptions in the range λ=510–540 nm. Complex 1 exhibited intense emission at λ=525 nm in 1 % DMSO/water solution (fluorescence quantum yield of 0.06). Nanosecond transient absorption spectral features indicated an enhanced population of the triplet excited state in di‐iodinated complex 2 . The generation of singlet oxygen by complex 2 upon exposure to visible light, as evidenced from experiments with 1,3‐diphenylisobenzofuran, is suitable for photodynamic therapy because of the remarkable photosensitizing ability. The complexes resulted in excellent photocytotoxicity in HaCaT cells (half maximal inhibitory concentration IC50≈3 μm , λ=400–700 nm, light dose=10 J cm?2), but they remained non‐toxic in the dark (IC50>100 μm ). Confocal microscopy images of 1 and Pt estimation from isolated mitochondria showed colocalization of the complexes in the mitochondria. Complex 2 displayed generation of reactive oxygen species induced by visible light, disruption of the mitochondrial membrane potential, and apoptosis.  相似文献   

2.
Novel indolocarbazole derivative 12‐(α‐L ‐arabinopyranosyl)indolo[2,3‐α]pyrrolo[3,4‐c]carbazole‐5,7‐dione (AIC) demonstrated high potency (at submicromolar concentrations) against the NCI panel of human tumor cell lines and transplanted tumors in vivo. In search of tentative targets for AIC, we found that the drug formed high affinity intercalative complexes with d(AT)20, d(GC)20 and calf thymus DNA (binding constants (1.6×106) M ?1Ka≤(3.3×106) M ?1). The drug intercalated preferentially into GC pairs of the duplex. Importantly, the concentrations at which AIC formed the intercalative complexes with DNA (C≤1 μM ) were identical to the concentrations that triggered p53‐dependent gene reporter transactivation, the replication block, the inhibition of topoisomerase I‐mediated DNA relaxation and death of HCT116 human colon carcinoma cells. We conclude that the formation of high affinity intercalative complexes with DNA is an important factor for anticancer efficacy of AIC.  相似文献   

3.
The present study describes the synthesis and anticancer activity of novel octahedral PtIV complexes with cyclohexyl functionalized ethylenediamine‐N,N′‐diacetate‐type ligands. Molecular mechanics calculations and density functional theory analysis revealed that s‐cis is the preferred geometry of these PtIV complexes with tetradentate‐coordinated (S,S)‐ethylenediamine‐N,N′‐di‐2‐(3‐cyclohexyl)propanoate. The viability of cancer cell lines (U251 human glioma, C6 rat glioma, L929 mouse fibrosarcoma, and B16 human melanoma) was assessed by measuring mitochondrial dehydrogenase activity and lactate dehydrogenase release. Cell‐cycle distribution, oxidative stress, caspase activation, and induction of autophagy were analyzed by flow cytometry using appropriate fluorescent reporter dyes. The cytotoxic activity of novel PtIV complexes against various cancer cell lines (IC50 range: 1.9–8.7 μM ) was higher than that of cisplatin (IC50 range: 10.9–67.0 μM ) and proceeded through completely different mechanisms. Cisplatin induced caspase‐dependent apoptosis associated with the cytoprotective autophagic response. In contrast, the new PtIV complexes caused rapid, caspase‐independent, oxidative stress‐mediated non‐apoptotic cell death characterized by massive cytoplasmic vacuolization, cell membrane damage, and the absence of protective autophagy.  相似文献   

4.
A series of novel fluorine‐containing cyclooxygenase‐2 (COX‐2) inhibitors was designed and synthesized based on the previously reported fluorescent COX‐2 imaging agent celecoxib–NBD ( 3 ; NBD=7‐nitrobenzofurazan). In vitro COX‐1/COX‐2 inhibitory data show that N‐(4‐fluorobenzyl)‐4‐(5‐p‐tolyl‐3‐trifluoromethylpyrazol‐1‐yl)benzenesulfonamide ( 5 ; IC50=0.36 μM , SI>277) and N‐fluoromethyl‐4‐(5‐p‐tolyl‐3‐trifluoromethylpyrazol‐1‐yl)benzenesulfonamide ( 6 ; IC50=0.24 μM , SI>416) are potent and selective COX‐2 inhibitors. Compound 5 was selected for radiolabeling with the short‐lived positron emitter fluorine‐18 (18F) and evaluated as a positron emission tomography (PET) imaging agent. Radiotracer [18F] 5 was analyzed in vitro and in vivo using human colorectal cancer model HCA‐7. Although radiotracer uptake into COX‐2‐expressing HCA‐7 cells was high, no evidence for COX‐2‐specific binding was found. Radiotracer uptake into HCA‐7 tumors in vivo was low and similar to that of muscle, used as reference tissue.  相似文献   

5.
Two silver(I) complexes—{[Ag(4‐pmOpe)]NO3}n and [Ag(2‐bimOpe)2]NO3—and three copper(II) complexes—[Cu4Cl6O(2‐bimOpe)4], [CuCl2(4‐pmOpe)2], and [CuCl2(2‐bis(pm)Ope]—were synthesized by reaction of silver(I) nitrate or copper(II) chloride with phosphate derivatives of pyridine and benzimidazole, namely diethyl (pyridin‐4‐ylmethyl)phosphate (4‐pmOpe), 1H‐benzimidazol‐2‐ylmethyl diethyl phosphate (2‐bimOpe), and ethyl bis(pyridin‐2‐ylmethyl)phosphate (2‐bis(pm)Ope). These compounds were characterized by 1H, 13C, and 31P NMR as well as IR spectroscopy, elemental analysis, and ESIMS spectrometry. Additionally, molecular and crystal structures of {[Ag(4‐pmOpe)]NO3}n and [Cu4Cl6O(2‐bimOpe)4] were determined by single‐crystal X‐ray diffraction analysis. The antimicrobial profiles of synthesized complexes and free ligands against test organisms from the ATCC and clinical sources were determined. Silver(I) complexes showed good antimicrobial activities against Candida albicans strains (MIC values of ~19 μM ). [Ag(2‐bimOpe)2]NO3 was particularly active against Pseudomonas aeruginosa and methicillin‐resistant Staphylococcus epidermidis, with MIC values of ~5 and ~10 μM , respectively. Neither copper(II) complexes nor the free ligands inhibited the growth of test organisms at concentrations below 500 μg mL?1.  相似文献   

6.
A series of 2,4,6‐trimethylbenzylidyne tungsten and molybdenum complexes was prepared from the tribromides mer‐[MesCMBr3(dme)] ( 9a , M=W; 9b , M=Mo, dme=1,2‐dimethoxyethane). Successive reaction of complexes 9 with lithium or potassium hexafluoro‐tert‐butoxide and lithium 1,3‐di‐tert‐butylimidazolin‐2‐imide, (Imt‐BuN)Li, afforded the imidazolin‐2‐iminato complexes [MesCM{OC(CF3)2Me}2(Imt‐BuN)] ( 10a , M=W; 10b , M=Mo), whereas the reactions of 9 with three equivalents of LiOSi(O‐t‐Bu)3 or KOC(CF3)3 gave [MesCM{OSi(O‐t‐Bu)3}3] ( 11a , M=W; 11b , M=Mo) and [MesCM{OC(CF3)3}3] ( 12a , M=W; 12b , M=Mo), respectively. For comparison, the benzylidyne complex [PhCMo{OSi(O‐t‐Bu)3}3] ( 7b ) was also prepared, and its molecular structure together with those of 10a , 10b , 11b , 12a and 12b were established by X‐ray diffraction analysis. Complexes 10 and 11 were employed as pre‐catalysts for the alkyne metathesis of the test substrate 3‐pentynyl benzyl ether ( 13 ) at low catalyst loadings (1 mol%) in the presence of molecular sieve (5 Å). Comparative studies of these 2,4,6‐trimethylbenzylidyne species (MesCM) with their benzylidyne analogues (PhCM) revealed that the increased steric bulk renders the former more stable and manageable in air in solid form for shorter periods of time, but at the expense of a slower initiation, which requires higher temperatures or longer reaction times.

  相似文献   


7.
Silsesquioxanes are employed as ligand backbones for the synthesis of novel phosphite compounds with 3,3′‐5,5′‐tetrakis(tert‐butyl)‐2,2′‐dioxa‐1,1′‐biphenyl substituents. Both mono‐ and bidentate phosphites are prepared in good yields. Two types of silsesquioxanes are employed as starting materials. The monophosphinite 1 and the monophosphite 2 are prepared from the thallium silsesquioxide derived from a completely condensed silsesquioxane framework (c‐C5H9)7Si7O12SiOTl. The diphosphite 3 is synthesized starting with the incompletely condensed monosilylated disilanol (c‐C5H9)7Si7O9(OSiMePh2)(OH)2. For monophosphite 2 , the corresponding trans‐[PtCl2( 2 )] complex 4 is characterized by NMR spectroscopy as well as by X‐ray crystallography, as the first example of a completely condensed oxo‐functionalized silsesquioxane framework. The coordination of the bidentate ligand 3 towards Pd, Mo and Rh is studied, both by NMR spectroscopy as well as by X‐ray crystallography. Various modes of coordination are shown to be possible. The molecular structures for the complexes trans‐[PdCl2( 3 )] ( 5 ), cis‐[Mo(CO)4( 3 )] ( 6 ) and the dinuclear complex [{Rh(μ‐Cl)(CO)}22‐ 3 )] ( 7 ) have been determined. In the rhodium‐catalyzed hydroformylation of 1‐octene high activities, with turnover frequencies of up to 6800 h−1, are obtained with these new nanosized phosphorus ligands.  相似文献   

8.
Two asymmetric alkylidene‐bridged dinuclear titanocenium complexes (CpTiCl2)25‐η5‐C9H6(CH2)nC5H4), 1 (n = 3) and 2 (n = 4) have been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligands C9H7(CH2)nC5H5 (n = 3, 4). Additionally, Ti(η55n‐BuC5H4C5H5)Cl2 (3) and Ti(η55n‐BuC9H6C5H5)Cl2 (4) were synthesized as corresponding mononuclear complexes. All complexes were characterized by 1H, 13C NMR, and IR spectroscopy. Homogenous ethylene polymerization catalyzation using those complexes has been conducted in the presence of methylaluminoxane (MAO). The influences of reaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results showed that the catalytic activities of both dinuclear titanocenes were higher than those of the corresponding mononuclear titanocenes. Although the two dinuclear complexes were different in only one [CH2] unit, the catalytic activity of 2 was about 50% higher than that of 1; however, the molecular weight of polyethylene (PE) obtained by 2 was lower than that obtained from 1. The molecular weight distribution of PE produced by these dinuclear complexes reached 6.9 and 7.3, respectively. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3317–3323, 2006  相似文献   

9.
Although the role of Bcl‐2 phosphorylation is still under debate, it has been identified in a resistance mechanism to BH3 mimetics, for example ABT‐737 and S1 . We identified an S1 analogue, S1‐16 , as a small‐molecule inhibitor of pBcl‐2. S1‐16 efficiently kills EEE‐Bcl‐2 (a T69E, S70E, and S87E mutant mimicking phosphorylation)‐expressing HL‐60 cells and high endogenously expressing pBcl‐2 cells, by disrupting EEE‐Bcl‐2 or native pBcl‐2 interactions with Bax and Bak, followed by apoptosis. In vitro binding assays showed that S1‐16 binds to the BH3 binding groove of EEE‐Bcl‐2 (Kd=0.38 μM by ITC; IC50=0.16 μM by ELISA), as well as nonphosphorylated Bcl‐2 (npBcl‐2; Kd=0.38 μM ; IC50=0.12 μM ). However, ABT‐737 and S1 had much weaker affinities to EEE‐Bcl‐2 (IC50=1.43 and >10 μM , respectively), compared with npBcl‐2 (IC50=0.011 and 0.74 μM , respectively). The allosteric effect on BH3 binding groove by Bcl‐2 phosphorylation in the loop region was illustrated for the first time.  相似文献   

10.
The complex [RhCl(NH2(CH2)12CH3)3] was tested for the semihydrogenation of 1‐heptyne in homogeneous and heterogeneous conditions. γ‐Al2O3 and two different commercial activated carbons (RX‐3 EXTRA and GF‐45) were used as supports. The results were, then, compared with those previously reported for the [PdCl2(NH2(CH2)12CH3)2] complex supported or unsupported, and with the results obtained with the classic Lindlar catalyst. The complex was characterised by FTIR and elemental analysis. The pure species and the supported one were also characterised by X‐ray photoelectron spectroscopy. Results determined by the latter technique suggest that the active species is the complex itself, which is stable under the reaction conditions. The supported rhodium tetra‐coordinated complex shows higher activity and selectivity than the same complex unsupported, and also than the classic Lindlar catalyst. Moreover, among the rhodium‐supported complexes the one immobilised on RX‐3 EXTRA has a better performance than that heterogenised on GF‐45, and this one has a better activity and selectivity than the γ‐Al2O3 anchored complex. Our results also show that under the same operational conditions (temperature, hydrogen pressure and metal/substrate weight ratio) the rhodium complex, unsupported or supported, has a better performance than the corresponding palladium complex. Copyright © 2004 Society of Chemical Industry  相似文献   

11.
A series of 1‐methyl‐1H‐indole–pyrazoline hybrids were designed, synthesized, and biologically evaluated as potential tubulin polymerization inhibitors. Among them, compound e19 [5‐(5‐bromo‐1‐methyl‐1H‐indol‐3‐yl)‐3‐(3,4,5‐trimethoxyphenyl)‐4,5‐dihydro‐1H‐pyrazole‐1‐carboxamide] showed the most potent inhibitory effect on tubulin assembly (IC50=2.12 μm ) and in vitro growth inhibitory activity against a panel of four human cancer cell lines (IC50 values of 0.21–0.31 μm ). Further studies confirmed that compound e19 can induce HeLa cell apoptosis, cause cell‐cycle arrest in G2/M phase, and disrupt the cellular microtubule network. These studies, along with molecular docking and 3D‐QSAR modeling, provide an important basis for further optimization of compound e19 as a potential anticancer agent.  相似文献   

12.
A group of cyclooxygenase‐2 (COX‐2)‐specific fluorescent cancer biomarkers were synthesized by linking the anti‐inflammatory drugs ibuprofen, (S)‐naproxen, and celecoxib to the 7‐nitrobenzofurazan (NBD) fluorophore. In vitro COX‐1/COX‐2 inhibition studies indicated that all of these fluorescent conjugates are COX‐2 inhibitors (IC50 range: 0.19–23.0 μM ) with an appreciable COX‐2 selectivity index (SI≥4.3–444). In this study the celecoxib–NBD conjugate N‐(2‐((7‐nitrobenzo[c][1,2,5]oxadiazol‐4‐yl)amino)ethyl)‐4‐(5‐(p‐tolyl)‐3‐(trifluoromethyl)‐1H‐pyrazol‐1‐yl)benzenesulfonamide ( 14 ), which displayed the highest COX‐2 inhibitory potency and selectivity (COX‐2 IC50=0.19 μM ; SI=443.6), was identified as an impending COX‐2‐specific biomarker for the fluorescence imaging of cancer using a COX‐2‐expressing human colon cancer cell line (HCA‐7).  相似文献   

13.
Different Rh complex catalysts were compared for the hydrogenation of methyl sorbate and linoleate in the absence of solvents. At 100 C and 1 atm H2 the following complexes, RhCl(Ph3 P)3 (Ph= phenyl), [RhClNBD]2 (NBD=norbornadiene) and RhH(CO)(Ph3P)3, produced mainly methyltrans-2-hexenoate (34 to 56%). Their diene selectivity was not particularly high as they produced 14 to 41% methyl hexanoate. With RhCl(Ph3 P)3 constant ratios between rates of methyl sorbate disappearance and formation of methyltrans-2- andtrans-3-hexenoate indicate approximately the same activation energy for 1,2-addition of H2 on the Δ4 double bond of methyl sorbate and for 1,4-addition to this substrate. In the hydrogenation of methyl linoleate with RhCl(Ph3 P)3, the kinetic curves were simulated by a scheme in which 1,2-reduction was more than twice as important as 1,4-addition of H2 via conjugated diene intermediates. Although the complexes RhCl(CO)(Ph3 P)3 and [Rh(NBD)(diphos)]+PF6 (diphos=diphosphine) were inactive in the hydrogenation of methyl sorbate, they catalyzed the hydrogenation of methyl linoleate at 100 C and 1 atm. Catalyst inhibition apparently was caused by stronger complex formation with methyl sorbate than with the conjugated dienes formed from methyl linoleate.  相似文献   

14.
A series of imidazo[2,1‐b][1,3,4]thiadiazole‐linked oxindoles composed of an A, B, C and D ring system were synthesized and investigated for anti‐proliferative activity in various human cancer cell lines; test compounds were variously substituted at rings C and D. Among them, compounds 7 ((E)‐5‐fluoro‐3‐((6‐p‐tolyl‐2‐(3,4,5‐trimethoxyphenyl)‐imidazo[2,1‐b][1,3,4]thiadiazol‐5‐yl)methylene)indolin‐2‐one), 11 ((E)‐3‐((6‐p‐tolyl‐2‐(3,4,5‐trimethoxyphenyl)imidazo[2,1‐b][1,3,4]thiadiazol‐5‐yl)methylene)indolin‐2‐one), and 15 ((E)‐6‐chloro‐3‐((6‐phenyl‐2‐(3,4,5‐trimethoxyphenyl)imidazo[2,1‐b][1,3,4]thiadiazol‐5‐yl)methylene)indolin‐2‐one) exhibited potent anti‐proliferative activity. Treatment with these three compounds resulted in accumulation of cells in G2/M phase, inhibition of tubulin assembly, and increased cyclin‐B1 protein levels. Compound 7 displayed potent cytotoxicity, with an IC50 range of 1.1–1.6 μM , and inhibited tubulin polymerization with an IC50 value (0.15 μM ) lower than that of combretastatin A‐4 (1.16 μM ). Docking studies reveal that compounds 7 and 11 bind with αAsn101, βThr179, and βCys241 in the colchicine binding site of tubulin.  相似文献   

15.
In this article, 2,2′‐bis[4‐(4‐maleimidephen‐oxy)phenyl)]propane (BMPP) resin and N,N‐4,4′‐bismaleimidodiphenylmethyene (BDM) resin blends were modified by diallyl bisphenol A (DABPA). The effects of the mole concentration of BMPP on mechanical properties, fracture toughness, and heat resistance of the modified resins were investigated. Scanning electron microscopy was used to study the microstructure of the fractured modified resins. The introduction of BMPP resin improves the fracture toughness and impact strength of the cured resins, whose thermal stabilities are hardly affected. Dynamic mechanical analysis shows that the modified resins can maintain good mechanical properties at 270.0°C, and their glass transition temperatures (Tg) are above 280.0°C. When the mole ratio of BDM : BMPP is 2 : 1(Code 3), the cured resin performs excellent thermal stability and mechanical property. Its Tg is 298°C, and the Charpy impact strength is 20.46 KJ/m2. The plane strain critical stress intensity factor (KIC) is 1.21 MPa·m0.5 and the plane strain critical strain energy release rate (GIC) is 295.64 J/m2. Compared with that of BDM/DABPA system, the KIC and GIC values of Code 3 are improved by 34.07% and 68.10%, respectively, which show that the modified resin presented good fracture toughness. © 2014 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2014 , 131, 40395.  相似文献   

16.
We describe the 3D‐QSAR‐assisted design of an Aurora kinase A inhibitor with improved physicochemical properties, in vitro activity, and in vivo pharmacokinetic profiles over those of the initial lead. Three different 3D‐QSAR models were built and validated by using a set of 66 pyrazole (Model I) and furanopyrimidine (Model II) compounds with IC50 values toward Aurora kinase A ranging from 33 nM to 10.5 μM . The best 3D‐QSAR model, Model III, constructed with 24 training set compounds from both series, showed robustness (r2CV=0.54 and 0.52 for CoMFA and CoMSIA, respectively) and superior predictive capacity for 42 test set compounds (R2pred=0.52 and 0.67, CoMFA and CoMSIA). Superimposition of CoMFA and CoMSIA Model III over the crystal structure of Aurora kinase A suggests the potential to improve the activity of the ligands by decreasing the steric clash with Val147 and Leu139 and by increasing hydrophobic contact with Leu139 and Gly216 residues in the solvent‐exposed region of the enzyme. Based on these suggestions, the rational redesign of furanopyrimidine 24 (clog P=7.41; Aurora A IC50=43 nM ; HCT‐116 IC50=400 nM ) led to the identification of quinazoline 67 (clog P=5.28; Aurora A IC50=25 nM ; HCT‐116 IC50=23 nM ). Rat in vivo pharmacokinetic studies showed that 67 has better systemic exposure after i.v. administration than 24 , and holds potential for further development.  相似文献   

17.
A number of aza‐heterocyclic compounds, which share the 5,6‐dihydropyrrolo[2,1‐a]isoquinoline (DHPIQ) scaffold with members of the lamellarin alkaloid family, were synthesized and evaluated for their ability to reverse in vitro multidrug resistance in cancer cells through inhibition of P‐glycoprotein (P‐gp) and/or multidrug‐resistance‐associated protein 1. Most of the investigated DHPIQ compounds proved to be selective P‐gp modulators, and the most potent modulator, 8,9‐diethoxy‐1‐(3,4‐diethoxyphenyl)‐3‐(furan‐2‐yl)‐5,6‐dihydropyrrolo[2,1‐a]isoquinoline‐2‐carbaldehyde, attained sub‐micromolar inhibitory potency (IC50: 0.19 μm ). Schiff bases prepared by the condensation of some 1‐aryl‐DHPIQ aldehydes with p‐aminophenol also proved to be of some interest, and one of them, 4‐((1‐(4‐fluorophenyl)‐5,6‐dihydro‐8,9‐dimethoxypyrrolo[2,1‐a]isoquinolin‐2‐yl)methyleneamino)phenol, had an IC50 value of 1.01 μm . In drug combination assays in multidrug‐resistant cells, some DHPIQ compounds, at nontoxic concentrations, significantly increased the cytotoxicity of doxorubicin in a concentration‐dependent manner. Studies of structure–activity relationships and investigation of the chemical stability of Schiff bases provided physicochemical information useful for molecular optimization of lamellarin‐like cytotoxic drugs active toward chemoresistant tumors as well as nontoxic reversers of P‐gp‐mediated multidrug resistance in tumor cells.  相似文献   

18.
A series of yttrium trisalicylaldimine complexes formed in situ by the reaction of trialkyl complex [Y(CH2SiMe3)3(THF)2] (THF is tetrahydrofuran) with three equivalent salicylaldimines were used as initiators for the ring‐opening polymerization of ε‐caprolactone. Electronic and steric effects of the salicylaldimine ligand played important roles on the catalytic properties of the yttrium complexes. The yttrium trisalicylaldimine complex Y( L7 )3 ( L7 = (S)‐2,4‐di‐tert‐butyl‐6‐[(1‐phenylethylimino)methyl]phenol) most effectively initiated controlled ring‐opening polymerization of ε‐caprolactone to prepare poly(ε‐caprolactone)s with high molecular weights and moderate molecular weight distributions. Obtained by density functional theory calculations, the optimized geometries of the four different active centers with four salicylaldimine ligands explained the experimental results. Copyright © 2011 Society of Chemical Industry  相似文献   

19.
Herein we describe the design, multicomponent synthesis, and biological, molecular modeling and ADMET studies, as well as in vitro PAMPA‐blood–brain barrier (BBB) analysis of new tacrine–ferulic acid hybrids (TFAHs). We identified (E)‐3‐(hydroxy‐3‐methoxyphenyl)‐N‐{8[(7‐methoxy‐1,2,3,4‐tetrahydroacridin‐9‐yl)amino]octyl}‐N‐[2‐(naphthalen‐2‐ylamino)2‐oxoethyl]acrylamide (TFAH 10 n ) as a particularly interesting multipotent compound that shows moderate and completely selective inhibition of human butyrylcholinesterase (IC50=68.2 nM ), strong antioxidant activity (4.29 equiv trolox in an oxygen radical absorbance capacity (ORAC) assay), and good β‐amyloid (Aβ) anti‐aggregation properties (65.6 % at 1:1 ratio); moreover, it is able to permeate central nervous system (CNS) tissues, as determined by PAMPA‐BBB assay. Notably, even when tested at very high concentrations, TFAH 10 n easily surpasses the other TFAHs in hepatotoxicity profiling (59.4 % cell viability at 1000 μM ), affording good neuroprotection against toxic insults such as Aβ1–40, Aβ1–42, H2O2, and oligomycin A/rotenone on SH‐SY5Y cells, at 1 μM . The results reported herein support the development of new multipotent TFAH derivatives as potential drugs for the treatment of Alzheimer′s disease.  相似文献   

20.
In this article, the binding characteristics of the imprinted polymer P‐1[CoII(salen)] (salen: bis(2‐hydroxybenzaldehyde)ethylenediimine) to nitric oxide (NO) have been reported. P‐1[CoII(salen)] was characterized by Fourier transform infrared analysis, thermogravimetric analysis, and differential scanning calorimetry. Batch‐mode adsorption studies were carried out to investigate binding thermodynamics, kinetics, and selective recognition behavior of P‐1[CoII(salen)] to NO. The kinetics study indicates that binding of the polymer to NO fits the first‐order reaction kinetics with the rate constant k1 of 0.087 min?1. Langmuir and Freundlich equations were used to explain the equilibrium character of P‐1[CoII(salen)] binding to NO. The r2 and χ2 values suggest that total amount of NO bound by P‐1[CoII(salen)] can be best fitted by the Langmuir equation. The binding capacity (Bmax) of P‐1[CoII(salen)] was calculated to be 76.28 μmol/g, very close to the experimental value, 75 μmol/g. The thermodynamics and selectivity experiments showed that the affinity of P‐1[CoII(salen)] to NO was much higher than carbon dioxide (CO2) and oxygen (O2), suggesting that P‐1[CoII(salen)] is a promising functional material for NO storage and NO sensing. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  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