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.
Utilizing a new microwave-assisted method, CuCoFe2O4@Chitosan (Ch) was synthesized as a very strong, magnetically separable nano-adsorbent. The magnetic nanohybrid adsorbent was characterized by FESEM (Field emission scanning electron microscopy), EDS (energy dispersive X-ray), Mapping & Linescan, BET (Brunauer-Emmett-Teller), FTIR (Fourier-transform infrared spectroscopy), XRD (X-ray diffraction analysis), TGA (Thermogravimetric analysis), and VSM (Vibrating Sample Magnetometer) techniques. Then, the adsorption process of Tetracycline (TC) was investigated. The highest percentage of pollutant adsorption on the synthetic and real samples was recorded at an initial concentration of 5 mg/L, pH 3.5, contact time of 20 min, the dose of 0.4 g/L, and temperature of 25 °C, 93.07 %, and 67%, respectively. The TC adsorption process via the synthesized magnetic nanocomposite was consistent with the Freundlich isotherm model (R2 = 0.992) and pseudo-second-order kinetic (K2 = 0.267). The outcomes of thermodynamic analyses, which included entropy changes (ΔS = 10.122 J/mol.k), enthalpy changes (ΔH = ?1.975 kJ/mol), and the Gibbs negative free energy (ΔG = ?4.992 kJ/mol), revealed that the adsorption process was spontaneous, favorable, and exothermic. The good magnetic properties allow easy separation after the adsorption operation. Finally, the efficiency of the nano-adsorbent in the removal process was 82.16% after four adsorption–desorption cycles. Some advantages of this research are a fast and green method for synthesis of adsorbent, fast kinetic, and magnetic properties to easy separation.  相似文献   

2.
An efficient and short enantioselective synthesis of (S)‐ and (R)‐tolterodine acid isomers ( 7a – 7i ) was performed a 6‐methyl‐4‐phenylchroman‐2‐one intermediate from inexpensive and commercially available starting materials. A series of tolterodine acid hybrids 7 were synthesized and characterized by infrared, 1H NMR, 13C NMR, X‐ray diffraction, and mass spectral analysis followed by anticancer activity on human cancer cell lines including A549 and SKNSH. Our results revealed the final compounds exhibited moderate to potent activity against A549 and SKNSH. Compounds 7g and 7f were more cytotoxic than cisplatin against all tested two human cancer cell lines, with half maximal inhibitory concentration values of 13.2, 14.3, and 8.5 μM, respectively. In the present investigation, possible binding interaction of the target compounds with 3IVX protein, ligand conformations, including hydrogen bonds and the bond lengths, was analyzed. AutoDock 4.2 chemokine receptor has been investigated by molecular docking and was used to predict the affinity, activity, and binding orientation of ligand with the target protein and to analyze best conformations. Compound 7h exhibited more binding energy (ΔG = −5.52 kcal/mol) and dissociation constant (KI = 89.8 μM) with amino acids Glu 17 and Thr 87 interacting. Further studies are warranted to fully evaluate the analogues as the potential prodrugs with improved physiochemical properties.  相似文献   

3.
Vapour pressures of water over saturated solutions of magnesium, calcium, nickel and zinc acetates were determined as a function of temperature. The vapour pressures served to evaluate the water activities, osmotic coefficients and molar enthalpies of vaporization. Molar enthalpies of solution of magnesium acetate tetrahydrate,ΔsolHm (T =  294.71K ;m =  0.01 mol · kg  1)  =   (15.65  ±  0.97)kJ · mol  1; calcium acetate,ΔsolHm (T =  297.18K ;m =  0.01 mol · kg  1)  =   (28.15  ±  0.28)kJ · mol  1; zinc acetate dihydrate,ΔsolHm (T =  297.36K ;m =  0.01 mol · kg  1)  =   (22.49  ±  0.90)kJ · mol  1and lead acetate trihydrate,ΔsolHm (T =  297.36K ;m =  0.0086 mol · kg  1)  =  (22.46  ±  0.94)kJ · mol  1, were determined calorimetrically.  相似文献   

4.
Detailed NMR studies of aqueous solutions (pH 7) of γ-cyclodextrin (γCD) and the azo dye Congo Red (CR) show distinct, concentration-independent 1H NMR signals for different species. A very stable 1:1 pseudorotaxane (K 11=38,000±1100?M-1) is formed. In addition, a second complex corresponding to a 2:2 adduct (K 22=13±3?M?-1) is produced by dimerisation of the 1:1 species. The structure of the 1:1 pseudorotaxane involves fast motion of the γCD ring along the CR backbone, leaving the outer naphthalene rings free. This entity undergoes structural reorganisation and dimerises to form the 2:2 adducts. Variable-temperature spectra did not lead to coalescence and allowed for the calculation of K 11 and K 22 at each temperature and also of the corresponding thermodynamic parameters. Therefore, formation of the 1:1 complex is favourable (ΔG=-26.1±0.1?kJ/mol) and exothermic (ΔH=-21.7±1.0?kJ/mol), whereas formation of the 2:2 entity is also favourable (ΔG=-6.36±0.58?kJ/mol) but endothermic (ΔH=+43.3±8.7?kJ/mol). The corresponding values for entropy change are both positive (ΔS 11=+14.5±0.7?J/mol, ΔS 22=+166±33?J/mol). Isothermal titration calorimetry studies confirm the NMR findings. For the 1:1 complexation, the dependence of K upon the concentration is indicative of the dimerisation to form the 2:2 complex. When CR is in excess, aggregation processes involving 2:2 complexes and CR molecules are observed by NMR and calorimetry.  相似文献   

5.
The interaction of human serum albumin (HAS) with divalent nickel ion was studied by isothermal titration calorimetry (ITC) in 30 mM Tris buffer, pH 7.0. There is a set of eight identical and independent binding sites for nickel ions on the protein at the temperature of 300 K. A new calorimetric data analysis allows the determination of the complete set of thermodynamic parameters. The binding isotherm for nickel-HSA interaction is easily obtained by carrying out two different ITC experiments. In the first experiment, the enthalpy of binding for one mole of nickel ion to one mole of binding site on HSA (ΔH=−36.5 kJ) is obtained, and is used in a second experiment to determine the binding isotherm and to find the number of binding sites (g=8) and the equilibrium constant (K=0.57 μM−1).  相似文献   

6.
《Arabian Journal of Chemistry》2020,13(10):7544-7557
Activated carbon (AC) derived from gasified Glyricidia sepium woodchip (GGSWAC) was prepared using KOH and CO2 activation via microwave radiation technique to remove atenolol (ATN) from aqueous solution. The surface area (SBET) and total pore volume (TPV) of GGSWAC were 483.07 m2/g and 0.255 cm3, respectively. The n-BET model fits well with the isothermal data indicating a multilayer adsorption with the saturation capacity of 121, 143 and 163 mg/g at 30, 45 and 60 °C, respectively. The kinetic study showed that ATN adsorption followed Avrami model equation (R2  0.99). Based on the thermodynamic parameters, the adsorption of ATN onto GGSWAC was endothermic (ΔHS = 234.17 kJ/mol) in the first layer of adsorption and exothermic in the subsequent layer (ΔHL = −165.62 kJ/mol). The ATN adsorption was controlled by both diffusion and chemisorption. In continuous operation, the Thomas (R2 = 0.9822) and Yoon–Nelson (R2 = 0.9817) models successfully predicted the ATN adsorption.  相似文献   

7.
Solubilization of caffeic acid into the aqueous solution of cationic cetyltrimethlyammonium bromide (CTAB) has been studied by using differential spectroscopic and conductivity methods. The solubility of caffeic acid increases with increasing the CTAB concentrations. The solubilization constant of caffeic acid into CTAB (KX = 1.8 × 105), standard free energy (ΔG0P = ?30.0 kJ/mol), and relative solubility (St/S0 = 53.5) were estimated at room temperature from UV–visible data. The critical micellar concentration (CMC) of CTAB decreases linearly with caffeic acid concentration due to the presence of hydrophobic benzene moiety. The interaction of caffeic acid with CTAB has also been discussed. The solubilized caffeic acid was used as a reducing agent for the preparation of silver nanoparticles (AgNPs). The as-prepared AgNPs were used as an activator of persulphate. The generated reactive oxygen species (OH?) and reactive sulphur species SO4-?) were responsible for the degradation of xylenol orange dye in water.  相似文献   

8.
Propyl gallate [3,4,5-trihydroxybenzoic acid propyl ester; PG] exhibits an anti-growth effect in various cells. In this study, the anti-apoptotic effects of various caspase inhibitors were evaluated in PG-treated Calu-6 and A549 lung cancer cells in relation to reactive oxygen species (ROS) and glutathione (GSH) levels. Treatment with 800 μM PG inhibited the proliferation and induced the cell death of both Calu-6 and A549 cells at 24 h. Each inhibitor of pan-caspase, caspase-3, caspase-8, and caspase-9 reduced the number of dead and sub-G1 cells in both PG-treated cells at 24 h. PG increased ROS levels, including O2∙−, in both lung cancer cell lines at 24 h. Generally, caspase inhibitors appeared to decrease ROS levels in PG-treated lung cancer cells at 24 h and somewhat reduced O2∙− levels. PG augmented the number of GSH-depleted Calu-6 and A549 cells at 24 h. Caspase inhibitors did not affect the level of GSH depletion in PG-treated A549 cells but differently and partially altered the depletion level in PG-treated Calu-6 cells. In conclusion, PG exhibits an anti-proliferative effect in Calu-6 and A549 lung cancer cells and induced their cell death. PG-induced lung cancer death was accompanied by increases in ROS levels and GSH depletion. Therefore, the anti-apoptotic effects of caspase inhibitors were, at least in part, related to changes in ROS and GSH levels.  相似文献   

9.
Solubilities of l -glutamic acid, 3-nitrobenzoic acid, p -toluic acid, calcium-l -lactate, calcium gluconate, magnesium- dl -aspartate, and magnesium- l -lactate in water were determined in the temperature range 278 K to 343 K. The apparent molar enthalpies of solution at T =  298.15 K as derived from these solubilities areΔsolHm (l -glutamic acid,msat =  0.0565 mol · kg  1)  =  30.2 kJ · mol  1,ΔsolHm (3-nitrobenzoic acid, m =  0.0188 mol · kg  1)  =  28.1 kJ · mol  1, ΔsolHm( p - toluic acid, m =  0.00267 mol · kg  1)  =  23.9 kJ · mol  1,ΔsolHm (calcium- l -lactate tetrahydrate,m =  0.2902 mol · kg  1)  =  25.8 kJ · mol  1,ΔsolHm (calcium gluconate, m =  0.0806 mol · kg  1)  =  22.1 kJ · mol  1, ΔsolHm(magnesium-dl -aspartate tetrahydrate, m =  0.1469 mol · kg  1)  =  11.5 kJ · mol  1, andΔsolHm (magnesium- l -lactate trihydrate,m =  0.3462 mol · kg  1)  =  3.81 kJ · mol  1.  相似文献   

10.
Methylcellulose (MC) is the most common commercial cellulose ether and the most attractive biopolymer due to its cheap cost of biodegradability, biocompatibility, hydrophilicity, and lack of toxicity. In this study, CoFe2O4@MC/activated carbon (AC) was synthesized as a unique magnetic nano-adsorbent in the presence of MC biopolymer for Reactive Red 198 (RR198) dye removal. The nano-magnetic adsorbent was characterized by FESEM (Field emission scanning electron microscopy), EDS (Energy-dispersive X-ray spectroscopy), Mapping, Linescan, BET (Brunauer–Emmett–Teller), FTIR (Fourier Transform Infrared Spectroscopy), XRD (X-Ray Diffraction), and VSM (Vibrating-Sample magnetometer). For simple separation by external magnetic fields, the Ms value was 57.91 emu/g. According to XRD analysis, the nano-adsorbent maintains its crystal structure, with an average crystal size of 11 nm. The maximum removal efficiencies of RR198 for synthetic and real wastewater samples under optimal conditions (an initial concentration of 10 mg/L, pH 3, contact time of 10 min, nanocomposite dose of 1.5 g/L, and a temperature of 25 °C) were 92.2% and 78%, respectively. The adsorption experiments were fitted well with the Freundlich isotherm (R2 = 0.989) and pseudo-second-order kinetic (R2 = 0.995). The values of entropy changes (ΔS = 35.087 kJ/mol.k), enthalpy changes (ΔH = -9.862 kJ/mol), and negative Gibbs free energy changes (ΔG) showed that the adsorption process was exothermic. Finally, the reusability findings showed that after six recovery cycles, the efficiency decreased slightly (90.1%). In the end, it can be concluded that the prepared CoFe2O4@Methylcellulose/AC can be used as an efficient adsorbent for the removal of RR198 from an aqueous solution.  相似文献   

11.
A series of copoly(methoxy-thiocyanurate)s is prepared in good yield and purity, and fully characterised. Many of the resulting polymers, formed at room temperature using phase transfer catalysis, can be cast into films with good resilience and thermal stability (some examples suffer practically no mass loss when held isothermally at 190 °C and only display appreciable losses when held continuously at 225 °C). Char yields of 61–64% are achieved in nitrogen depending on backbone structure. Some problems were encountered with solubility, particularly with copolymers, which limited molecular weights analysis, but values of Mn = 7000–10,000 g mol−1 were obtained for the polycyanurate and polythiocyanurate homopolymers. DSC reveals polymerisation exotherms with maxima at 197–207 °C (ΔHp = 39–48 kJ/mol), which are believed to be due to isomerisation of the (activation energies span 172–205 kJ/mol), since X-ray powder diffraction measurements reveal no evidence of crystalline structure in the resulting product.  相似文献   

12.
White spot syndrome virus (WSSV) remains as one of the most dreadful pathogen of the shrimp aquaculture industry owing to its high virulence. The cumulative mortality reaches up to 100% within in 2–10 days in a shrimp farm. Currently, no chemotherapeutics are available to control WSSV. The viral envelope protein, VP28, located on the surface of the virus particle acts as a vital virulence factor in the initial phases of inherent WSSV infection in shrimp. Hence, inhibition of envelope protein VP28 could be a novel way to deal with infection by inhibiting its interaction in the endocytic pathway. In this direction, a timely attempt was made to recognize a potential drug candidate of marine origin against WSSV using VP28 as a target by employing in silico docking and molecular dynamic simulations. A virtual library of 388 marine bioactive compounds was extracted from reports published in Marine Drugs. The top ranking compounds from docking studies were chosen from the flexible docking based on the binding affinities (ΔGb). In addition, the MD simulation and binding free energy analysis were implemented to validate and capture intermolecular interactions. The results suggested that the two compounds obtained a negative binding free energy with −40.453 kJ/mol and −31.031 kJ/mol for compounds with IDs 30797199 and 144162 respectively. The RMSD curve indicated that 30797199 moves into the hydrophobic core, while the position of 144162 atoms changes abruptly during simulation and is mostly stabilized by water bridges. The shift in RMSD values of VP28 corresponding to ligand RMSD gives an insight into the ligand induced conformational changes in the protein. This study is first of its kind to elucidate the explicit binding of chemical inhibitor to WSSV major structural protein VP28.  相似文献   

13.
The diamide N,N,N,N′-tetraoctyldiglycolamide (TODGA) was synthesized and characterized. The prepared TODGA was applied for extraction of Ce(III) from nitric acid solutions. The equilibrium studies included the dependencies of cerium distribution ratio on nitric acid, TODGA, nitrate ion, hydrogen ion and cerous ion concentrations. Analysis of the results indicates that the main extracted species is Ce(TODGA)2(NO3)3HNO3. The capacity of Ce loading is approximately 45 mmol/L for 0.1 M solution of TODGA in n-hexane. Finally, the thermodynamic parameters were calculated: K (25 °C) = 3.8 × 103, ΔH = −36.7 ± 1.0 kJ/mol, ΔS = −54.6 ± 3.0 J/K mol, and ΔG = −20.4 ± 0.1 kJ/mol.  相似文献   

14.
Our previous experimental results have shown that ergosta‐4,6,8(14),22‐tetraen‐3‐one (ergone) is one of the main bioactive components of Polyporus umbellatus. The efficacy of ergone binding to human serum albumin (HSA) is critical for pharmacokinetic behavior of ergone. The interactions between ergone and HSA under simulative physiological conditions were investigated by the methods of fluorescence spectroscopy, absorption and circular dichroism spectroscopy. Fluorescence data revealed that the fluorescence quenching of HSA by ergone was the result of the formation of the ergone‐HSA complex. According to the modified Stern‐Volmer equation, the binding constants (Ka) between ergone and HSA were determined. The thermodynamic parameters, enthalpy change (ΔH) and entropy change (ΔS) for the reaction were calculated to be 0.989 kJ mol‐1 and 11.214 J mol‐1 K‐1, indicating that the hydrogen bonds and hydrophobic interactions played a dominant role in the binding of ergone to HSA. The conformational investigation showed that the presence of ergone decreased the α‐helical content of HSA and induced the slight unfolding of the polypeptides of protein. Furthermore, displacement experiments using warfarin and ibuprofen indicated that ergone could bind to site I of HSA, which was also in agreement with the results of the molecular modeling.  相似文献   

15.
We studied the inhibitory effects of trifluoroethanol (TFE) on the activity and conformation of tyrosinase. TFE increased the degree of secondary structure of tyrosinase, which directly resulted in enzyme inactivation. A reciprocal study showed that TFE inhibited tyrosinase in a slope-parabolic mixed-type inhibition manner (K I = 0.5 ± 0.096 M). Time-interval kinetic studies showed that the inhibition was best described as first order with biphasic processes. Intrinsic and 1-anilinonaphthalene-8-sulfonate-binding fluorescences were also measured to gain more insight into the supposed structural changes; these showed that TFE induced a conspicuous tertiary structural change in tyrosinase by exposing hydrophobic surfaces. We also predicted the tertiary structure of tyrosinase and simulated its docking with TFE. The docking simulation was successful with significant scores (binding energy for Autodock4 = −4.75 kcal/mol; for Dock6 = −23.07 kcal/mol) and suggested that the TRP173 residue was mainly responsible for the interaction with TFE. Our results provide insight into the structure of tyrosinase and allow us to describe a new inhibition strategy that works by inducing conformational changes rather than targeting the active site of the protein.  相似文献   

16.
In this paper, several rare earth [terbium(III), ytterbium(III) and yttrium(III)] complexes containing 2,9-dimethyl-1,10-phenanthroline (Me2Phen) were successfully synthesized and characterized by means of elemental analysis (CHN), infrared spectroscopy (FT-IR), UV–vis absorption spectroscopy and 1HNMR. To explore the potential medicinal value of these complexes (MMe2Phen), their binding interactions with human serum albumin (HSA) were investigated through UV–vis and fluorescence spectroscopies and also molecular docking examinations. The thermodynamic parameters, binding forces and Förster resonance distance between these complexes and Trp-214 of HSA were estimated from the analysis of fluorescence measurements. The values of estimated binding constants (Kb) ranging for the formation of MMe2Phen:HSA complex were in the order of 105 M?1. The thermodynamic parameters determined by van’t Hoff analysis of KbH°?<?0 and ΔS°?<?0) clearly indicate the major rules of hydrogen bonds and van der Waals interactions in the formation process of MMe2Phen:HSA. The values of Stern–Volmer constant and the evaluation of dynamic quenching constant at various temperatures provided good evidences for static quenching mechanism. Furthermore, the results of molecular docking calculation and competitive binding experiments represent the binding of these complexes to site 3 of HSA located in subdomain IB, containing both polar and apolar residues. The consistency of computational and experimental results, according to the binding sites and the order of binding affinities (TbMe2Phen?>?YbMe2Phen?>?YMe2Phen), supports the accuracy of docking calculation.  相似文献   

17.
The aim of this study is to shed more light on the formation of mullite and the kinetics of mullitization from sol-gel synthesized precursors. Tetraethylorthosilicate (TEOS) and aluminum nitrate nonahydrate (ANN) were used, as a source of silica and alumina, respectively, for the synthesis of homogenous mullite precursor powder. The mullitization process was characterized by thermogravimetry (TG), differential thermal analysis (DTA), thermodilatometric analysis (TDA), and x-ray powder diffraction (XRD) techniques. It was found that mullite started to crystalize at temperatures of 1050, 1200, and 1241 °C as determined by XRD, DTA, and TDA, respectively?. Mullite crystallization kinetics was thoroughly investigated under isothermal and non-isothermal conditions using DTA. The activation energy for mullite formation was calculated, for different crystallization fractions, following the Freidman, Kissinger, Boswell, and Ozawa methods. The average values were found to be 1282.92, 1324.30, 1336.93, and 1283.09 kJ/mol, respectively. The kinetic parameters and the crystallization mechanism were determined and the results were compared with those available in the literature. The Sestak Berggren SB(m,n) model was found to be the most suitable for the determination of mullite crystallization mechanism. The calculated average values of the Gibbs free energy (ΔG#), enthalpy (ΔH#), and entropy (ΔS#) for mullite formation, at different heating rates, were 433.98 kJ/mol, 1294.20 kJ/mol, and 566.23 J/mol.K, respectively.  相似文献   

18.
19.
The interaction of ginkgolic acid (15:1, GA) with human serum albumin (HSA) was investigated by FT–IR, CD and fluorescence spectroscopic methods as well as molecular modeling. FT–IR and CD spectroscopic showed that complexation with the drug alters the protein’s conformation by a major reduction of α-helix from 54 % (free HSA) to 46–31 % (drug–complex), inducing a partial protein destabilization. Fluorescence emission spectra demonstrated that the fluorescence quenching of HSA by GA was by a static quenching process with binding constants on the order of 105 L·mol?1. The thermodynamic parameters (ΔH = ?28.26 kJ·mol?1, ΔS = 11.55 J·mol?1·K?1) indicate that hydrophobic forces play a leading role in the formation of the GA–HSA complex. The ratio of GA and HSA in the complex is 1:1 and the binding distance between them was calculated as 2.2 nm based on the Förster theory, which indicates that the energy transfer from the tryptophan residue in HSA to GA occurs with high probability. On the other hand, molecular docking studies reveal that GA binds to Site II of HSA (sub-domain IIIA), and it also shows that several amino acids participate in drug–protein complexation, which is stabilized by H-bonding.  相似文献   

20.
Tempol (4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl) is a stable, cell-permeable redox-cycling nitroxide water-soluble superoxide dismutase (SOD) mimetic agent. However, little is known about its cytotoxic effects on lung-related cells. Thus, the present study investigated the effects of Tempol on cell growth and death as well as changes in reactive oxygen species (ROS) and glutathione (GSH) levels in Calu-6 and A549 lung cancer cells, normal lung WI-38 VA-13 cells, and primary pulmonary fibroblast cells. Results showed that Tempol (0.5~4 mM) dose-dependently inhibited the growth of lung cancer and normal cells with an IC50 of approximately 1~2 mM at 48 h. Tempol induced apoptosis in lung cells with loss of mitochondrial membrane potential (MMP; ∆Ψm) and activation of caspase-3. There was no significant difference in susceptibility to Tempol between lung cancer and normal cells. Z-VAD, a pan-caspase inhibitor, significantly decreased the number of annexin V-positive cells in Tempol-treated Calu-6, A549, and WI-38 VA-13 cells. A 2 mM concentration of Tempol increased ROS levels, including O2•− in A549 and WI-38 VA-13 cells after 48 h, and specifically increased O2•− levels in Calu-6 cells. In addition, Tempol increased the number of GSH-depleted cells in Calu-6, A549, and WI-38 VA-13 cells at 48 h. Z-VAD partially downregulated O2•− levels and GSH depletion in Tempol-treated these cells. In conclusion, treatment with Tempol inhibited the growth of both lung cancer and normal cells via apoptosis and/or necrosis, which was correlated with increased O2•− levels and GSH depletion.  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