首页 | 官方网站   微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 359 毫秒
1.
Removal of vanadium from vanadyl (IV) meso-tetra-phenylporphyrin (VO-MTPP) and vanadyl (IV) octa-ethylporphyrin (VO-OEP) and metals removal from Ayacucho Venezuelan crude oil samples were performed using electrochemical techniques. According to cyclic voltammograms, a potential of − 2.3 V vs. Ag/AgNO3 (0.1 M), LiClO4 (0.1 M) in acetonitrile, was chosen for running electrolysis at platinum (Pt), graphite (G) and glassy carbon (GC) electrodes. Qualitative analysis was done by cyclic voltammetry (CV) and atomic absorption spectroscopy (AAS) while quantification was performed by AAS and inductively coupled plasma/atomic emission spectrometry (ICP/AES). Three stages (best results) of the study are reported: (a) when using commercial porphyrins; the best conditions for electrolysis were: tetrahydrofuran (THF)/20% methanol (MeOH)/1% perchloric acid (HClO4) on GC, producing 84% hydro-demetalation (HDM) for VO-MTPP, and 78% HDM for VO-OEP; (b) when using extracts of crude oil; demetalation percentages, after 90 min of electrolysis, on graphite, and after 120 min of electrolysis on platinum, were 66.44% and 64.10% HDM respectively, no discrimination of metals under these conditions (c) from electrolysis in whole crude oil, quantitative analysis gave: vanadium (V) 31.2 mg/kg, iron (Fe) 65, 0 mg/kg and nickel (Ni) 6, 3 mg/kg, with charge efficiencies of V (7.5%), Fe (79.6%) and Ni (8.2%).  相似文献   

2.
Three sulfur-containing methylphosphonates, dibutyl propylthiomethylphosphonate, (C4H8O)2P(0)CH2SC3H7,(DBPTMP),dibutyl propylsulfinylmethylphosphonate, (C4H90)2P(0)CH2S(0)C3H7, (DBPSiMP) and dibutyl propylsulf onylmethylphosphonate, (C4H8O)2P(0)CH2S(0) C3H7 (DBPSiMP) were synthesized. The extraction of gold(III), palladium(II) and platinum(IV) from HCl media with the synthesized compounds and dibutyl 2-keto-3-methylbutyl-phosphonate, (C4H8O)2P(0)CH2C (0) CH(CH3) CH3, (DBKPP) was investigated at various mixing times and acidities. DBPTMP and DBPSiMP were found to provide nearly quantitative extraction of Au(III) and Pd(II) with short equilibration times for Pd(II) extraction. DBKPP was selective for Au(TII) over Pt(IV) and Pd(II) at high acidity. The behavior of these extractants was also examined in comparison with the corresponding monofunctional extractants, dibutyl butylphosphonate(DB‘BP’), tributyl-phosphate(TBP), dioctylsulphide(DOS) and dioctyl-sulphoxide(DOSO).  相似文献   

3.
Conductimetric measurements have been performed on solutions of the following salts in N,N-dimethylthioformamide (DMTF) at 25°C:- triisoamyl-n-butylammonium tetraphenylborate, triisoamyl-n-butylammonium iodide, triisoamyl-n-butylammonium perchlorate, tetra-n-butylammonium tetraphenylborate, tetra-n-butylammonium iodide, tetra-n-butylammonium perchlorate, bis(DMTF)copper(I)chloride, bis(DMTF)copper(I) bromide, bis(DMTF)copper(I) iodide, tetrakis(DMTF)copper(I) perchlorate, bis(DMTF)silver(I) chloride, bis(DMTF)silver(I) bromide and (DMTF)silver(I) iodide. The data have been analysed with the Pitts and the Fuoss—Justice conductivity equations and single ion conductivities have been calculated based on the assumption of equal limiting anion and cation conductivities for triisoamyl-n-butylammonium tetraphenylborate.  相似文献   

4.
In this study, the effect of n-3 fatty acids (FA) [α-linolenic acid (ALA) and eicosapentaenoic acid (EPA) + docosahexaenoic acid (DHA)] on the intestinal bile acid (BA) uptake, hepatic BA synthesis, and enterohepatic bile acid transporters (BAT) was assessed in young and aged dyslipidemic rats. Dyslipidemia was induced in young and aged rats by feeding a high-fat (HF) diet. Experimental groups received diets containing canola oil (HF + CNO) and fish oil (HF + FO) as a source of ALA and EPA + DHA, respectively. After 60 days of feeding, intestinal BA uptake and expression of apical sodium-dependent bile acid transporter (Asbt), organic solute transporter-alpha/beta (Osta/b) messenger RNA (mRNA), and hepatic expression of Na+ taurocholate cotransporting polypeptide (Ntcp), bile salt export pump (Bsep), cholesterol 7-α hydroxylase A1 (Cyp7a1), Farnesoid X receptor (Fxr), small heterodimer partner-1 (Shp), liver receptor homolog-1 (Lrh-1), and hepatic nuclear factor-4 alpha (Hnf4a) mRNA were measured. Hepatic 3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) reductase activity and total BA in serum, liver, and feces were assessed. The dyslipidemic HF group had: (1) increased intestinal BA uptake and Asbt and Osta/b mRNA expression, (2) increased BA in serum, (3) decreased hepatic expression of Ntcp, Bsep, and Cyp7a1 mRNA, (4) increased activity of HMG-CoA reductase, (5) increased hepatic expression of Fxr and Shp mRNA, (6) decreased hepatic expression of Lrh-1 and Hnf4a mRNA, and (7) decreased BA in feces, when compared to control, HF + CNO, and HF + FO groups. Immunostaining revealed increased expression of intestinal Asbt and hepatic Ntcp protein in the HF group when compared to control, HF + CNO, and HF + FO groups. n-3 FA abrogated dyslipidemia-induced changes in the intestinal uptake, hepatic synthesis, and enterohepatic transporters of BA in both young and aged rats. EPA + DHA was more effective than ALA in modulating dyslipidemia-induced changes.  相似文献   

5.
The reaction of 3-mercaptopropan-l,2-diol with methyl 10-undecenoate (Ib) yielded three products, methyl 11-(3′-mercaptopropan-l′,2′-diacetoxy) undecanoate (II, 48.9%), methyl 11-(1-oxapropan-2′-acetoxy-3′-mercaptoacetyl) undecanoate (III, 27.4%) and methyl 10-(3′-mercaptopropan-1′-acetoxy-2′-ol) undecanoate (IV, 23.0%) along with hydrolyzed starting material (Ia, 5.4%). The same reaction with methyl 9-octadecenoate (Vb) gave an isomeric product, 9(10)-(3′-mercaptopropan-l′-acetoxy-2′-ol) octadecanoic acid (VI, 78.5%) and oleic acid (Va, 21.4%). Reaction withtrans-2-octadecenoic acid (VII) afforded 2(3)-(3′-mercaptopropan-l′-acetoxy-2′-ol) octadecanoic acid (VIII, 85.7%).  相似文献   

6.
In this study, the properties of blue organic light-emitting diodes (OLEDs), employing quantum well-like structure (QWS) that includes four different blue emissive materials of 4,4′-bis(2,2′-diphenylyinyl)-1,1′-biphenyl (DPVBi), 9,10-di(naphth-2-yl)anthracene (ADN), 2-(N,N-diphenyl-amino)-6-[4-(N,N-diphenyl amine)styryl]naphthalene (DPASN), and bis(2-methyl-8-quinolinolate)-4-(phenyl phenolato) aluminum (BAlq), were investigated. Conventional QWS blue OLEDs composed of multiple emissive layers and charge blocking layer with lower highest occupied molecular orbital (HOMO)-lowest unoccupied molecular orbital (LUMO) energy level, and devices with triple emissive layers for more significant hole-electron recombination and a wider region for exciton generation were designed. The properties of triple emissive layered blue OLEDs with the structure of indium tin oxide (ITO) /N,N′-diphenyl-N,N′-bis(1-naphthyl-phenyl)-(1,1′-biphenyl)-4,4′-diamine (NPB) (700 Ǻ)/X (100 Ǻ)/BAlq (100 Ǻ)/X (100 Ǻ)/4,7-diphenyl-1,10-phenanthroline (Bphen) (300 Ǻ)/lithium quinolate (Liq) (20 Ǻ)/aluminum (Al) (1,200 Ǻ) (X = DPVBi, ADN, DPASN) were examined. HOMO-LUMO energy levels of DPVBi, ADN, DPASN, and BAlq are 2.8 to 5.9, 2.6 to 5.6, 2.3 to 5.2, and 2.9 to 5.9 eV, respectively. The OLEDs with DPASN/BAlq/DPASN QWS with maximum luminous efficiency of 5.32 cd/A was achieved at 3.5 V.  相似文献   

7.
The crystallization behavior of the stereoblock copolymer of substituted and non-substituted poly(lactide)s, i.e., poly(d-2-hydroxybutyrate) and poly(l-lactide) chains having the opposite configurations [P(D-2HB)-b-PLLA] and the reference block copolymer of poly(d-2-hydroxybutyrate) and poly(d-lactide) chains with the identical configurations [P(D-2HB)-b-PDLA] was investigated. At the crystallizable temperature range of 60-160 °C, the crystallized P(D-2HB)-b-PLLA contained solely the hetero-stereocomplex crystallites as a crystalline species, without formation of poly(d-2-hydroxybutyrate) or poly(l-lactide) homo-crystallites, in contrast with their polymer blends. On the other hand, at the crystallizable temperature range of 60-140 °C, the crystallized P(D-2HB)-b-PDLA had only PDLA homo-crystallites as crystalline species, reflecting no co-crystallites formation between poly(d-2-hydroxybutyrate) and poly(d-lactide) chains having the same configurations. The equilibrium melting temperature of hetero-stereocomplex crystallites in P(D-2HB)-b-PLLA was 189.0 °C, which was higher than 171.3 °C of PDLA homo-crystallites in P(D-2HB)-b-PDLA. Although the final crystallinity of P(D-2HB)-b-PLLA was higher than those of P(D-2HB)-b-PDLA, the spherulite growth rate of P(D-2HB)-b-PLLA was lower.The regime analysis indicated unusual nucleation mechanism of P(D-2HB)-b-PLLA.  相似文献   

8.
Knoevenagel condensation of 3,4‐dichloro‐ and 2,6‐dichlorophenylacetonitriles gave a library of dichlorophenylacrylonitriles. Our leads (Z)‐2‐(3,4‐dichlorophenyl)‐3‐(1H‐pyrrol‐2‐yl)acrylonitrile ( 5 ) and (Z)‐2‐(3,4‐dichlorophenyl)‐3‐(4‐nitrophenyl)acrylonitrile ( 6 ) displayed 0.56±0.03 and 0.127±0.04 μm growth inhibition (GI50) and 260‐fold selectivity for the MCF‐7 breast cancer cell line. A 2,6‐dichlorophenyl moiety saw a 10‐fold decrease in potency; additional nitrogen moieties (‐NO2) enhanced activity (Z)‐2‐(2,6‐dichloro‐3‐nitrophenyl)‐3‐(2‐nitrophenyl)acrylonitrile ( 26 ) and (Z)‐2‐(2,6‐dichloro‐3‐nitrophenyl)‐3‐(3‐nitrophenyl)acrylonitrile ( 27 ), with the corresponding ‐NH2 analogues (Z)‐2‐(3‐amino‐2,6‐dichlorophenyl)‐3‐(2‐aminophenyl)acrylonitrile ( 29 ) and (Z)‐2‐(3‐amino‐2,6‐dichlorophenyl)‐3‐(3‐aminophenyl)acrylonitrile ( 30 ) being more potent. Despite this, both 29 (2.8±0.03 μm ) and 30 (2.8±0.03 μm ) were found to be 10‐fold less cytotoxic than 6 . A bromine moiety effected a 3‐fold enhancement in solubility with (Z)‐3‐(5‐bromo‐1H‐pyrrol‐2‐yl)‐2‐(3,4‐dichlorophenyl)acrylonitrile 18 relative to 5 at 211 μg mL?1. Modeling‐guided synthesis saw the introduction of 4‐aminophenyl substituents (Z)‐3‐(4‐aminophenyl)‐2‐(3,4‐dichlorophenyl)acrylonitrile ( 35 ) and (Z)‐N‐(4‐(2‐cyano‐2‐(3,4‐dichlorophenyl)vinyl)phenyl)acetamide ( 38 ), with respective GI50 values of 0.030±0.014 and 0.034±0.01 μm . Other analogues such as 35 and 36 were found to have sub‐micromolar potency against our panel of cancer cell lines (HT29, colon; U87 and SJ‐G2, glioblastoma; A2780, ovarian; H460, lung; A431, skin; Du145, prostate; BE2‐C, neuroblastoma; MIA, pancreas; and SMA, murine glioblastoma), except compound 38 against the U87 cell line. A more extensive evaluation of 38 ((Z)‐N‐(4‐(2‐cyano‐2‐(3,4‐dichlorophenyl)vinyl)phenyl)acetamide) in a panel of drug‐resistant breast carcinoma cell lines showed 10–206 nm potency against MDAMB468, T47D, ZR‐75‐1, SKBR3, and BT474. Molecular Operating Environment docking scores showed a good correlation between predicted binding efficiencies and observed MCF‐7 cytotoxicity. This supports the use of this model in the development of breast‐cancer‐specific drugs.  相似文献   

9.
The interactions between the water-soluble polyelectrolytes poly(acrylic acid) (PAA) and poly(vinyl sulfonic acid) (PVS), and Cu(II) and Ni(II) are studied by the liquid-phase polymer-based retention (LPR) technique. Assuming a Ni(II)-PVS interaction of electrostatic nature, the nature of the Ni(II)-PAA interaction is found to be electrostatic, while Cu(II)-PAA interactions imply the formation of coordinative bonds. The charge related formation constants for the systems Ni(II)-PAA, Ni(II)-PVS, and Cu(II)-PVS are found to be 57.57×102, 43.4×102, and 60.5×102 M−1, respectively in a 0.010 M NaNO3 aqueous solution at pH 5, and 1.4×102 for both systems containing Ni(II) and 1.3×102 M−1 for the system Cu(II)-PVS in a 0.10 M NaNO3 aqueous solution at pH 5.  相似文献   

10.
The effect of side chain length of π-conjugated poly(9,9-dialkylfluorene)s has been studied in semi-dilute (10 mg/mL) toluene solutions using small-angle neutron scattering (SANS) and 1H and 2H NMR spectroscopies. Under these conditions, SANS data indicate that poly(9,9-dinonylfluorene) and poly(9,9-dioctylfluorene) are dissolved down to the molecule level and appear as elongated one-dimensional chains (length >20-30 nm). In contrast, the shorter side chain polymers exhibit a self-association so that poly(9,9-diheptylfluorene) forms thin sheet-like (∼1 nm) and poly(9,9-dihexylfluorene) thin (∼1 nm) and thick sheet-like (>6 nm) aggregates. 1H NMR data, together with the density functional theory (DFT) calculations, however, show that this occurs without changes in the conformation of the polymer backbone.  相似文献   

11.
Various bis(silyl)ethenyl groups were attached to the aromatic ring of poly(α-methylstyrene) via Peterson olefination reaction of (RMe2Si)3CLi (R = H, Me and Ph) with formylated poly(α-methylstyrene) (Pα-MS-CHO) to give poly(α-methylstyrene)-co-[2,2-bis(silyl)ethenyl(α-methylstyrene)] as functionalized poly(α-methylstyrene) (Pα-MS-SiMe2H, Pα-MS-SiMe3 and Pα-MS-SiMe2Ph). The trimethylsilyl groups of Pα-MS-SiMe3 have been converted to 2,2-dibromoethenyl and epoxybis(silanes) groups via NBS-based bromodesilylation and MCPBA-based epoxidation respectively. The thermal degradation behaviors of the polymers were studied using thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC).  相似文献   

12.
The pure crystalline cerium oxide (CeO2) nanoparticles were synthesized using optimized content of Ce(NO3)3. 6H2O with varying concentrations of sodium hydroxide (NaOH) (0.5, 1, 1.5, and 2 M) as a precipitation agent in presence of 2.5 wt% poly(vinylpyrrolidone) PVP. All the samples are prepared via the modified coprecipitation technique. The synthesized materials have been analyzed using X-ray diffraction (XRD), Fourier transform infrared (FT-IR), laser Raman, high-resolution scanning electron microscope (HR-SEM), and photo luminescence (PL) analyses. The optimized sample was identified with the help of the above studies that could be analyzed through transverse electron microscopy (TEM) and X-ray photoelectron spectroscopy (XPS) studies. The cubic structure with the Fm-3 m space group has been confirmed through XRD (JCPDS: 81-0792) and Raman analyses. The FT-IR and energy dispersive X-ray spectroscopy (EDX) analyses ascertain the occurrence of Ce and O species. The as-prepared CeO2 filler (0, 3, 6, 9, and 12 wt%) is dispersed through the optimized polymer electrolyte Poly (styrene-co-methyl methacrylate) P(S-MMA) (27 wt%)–lithium perchloride (LiClO4) (8 wt%)–ethylene carbonate + propylene carbonate (EC + PC) (1;1 of 65 wt%) complex system using solution casting technique. P(S-MMA) (27 wt%)–LiClO4 (8 wt%)–EC + PC (1;1 of 65 wt%)–6 wt% of CeO2 shows the high ionic conductivity 8.13 × 10−4 S cm−1.  相似文献   

13.
《分离科学与技术》2012,47(5):1263-1269
Abstract

Solvent extraction behavior of Th(IV) and U(VI) and some lanthanides [Ce(III), Nd(III), Eu(III), Tb(III), and Yb(III)] from thiocyanate medium into sulfoxides and/or 2-thenoyltrifluoroacetone has been studied. The actinides are found to be favorably extracted by both the extractants. The alkyl sulfoxides extract Th(IV) and U(VI) as Th(SCN)4·3DPSO, Th(SCN)4·3DOSO, UO2 (SCN)2·2DPSO, and UO2(SCN)2·3DOSO. The chelate extracts the metals in the following order: U(VI) < Th(IV) < Eu(IV) < Tb(III) < Tb(III) < Ce (III) or Nd(III).  相似文献   

14.
The crystal structure and molecular conformation of 2-cyano-4-bromo-4′-N,N-diethylaminoazobenzene (C17H17N4Br, mol. wt. = 357·2 a.m.u) has been determined from X-ray diffraction data; triclinic, P1 (No. 2), a = 10·132(11) Å, b = 12·216(16) Å, c = 6·966(11) Å, α = 104·21(9)°, β = 92·67(12)°, γ = 97·22(7)°, V = 826·5(9) Å3, Z = 2, Dc = 1·436 g cm?3, F(000) = 378, λ(MoKα) = 0·71069 Å, μ(MoKα) = 26·0 cm?1. The structure was solved by the multiple solution direct method and refined by full-matrix least-squares to R = 0·059 for 1538 independent observed reflections. The azobenzene skeleton is planar to within 0·06 Å. Most significant bonding data are: NN, 1·290(8) Å; BrC, 1·866(6) Å; mean CN (azo) 1·380(8) Å; NNC, 113·6(4) and 115·3(4)°; NCC (cis relative to NN) 125·9(4)° and 126·7(4)°; NCC (trans) 116·8°(5)° and 116·1(4)°.  相似文献   

15.
Transition metal complexes of Mn(II), Fe(III), Cr(III) and Zn(II) metal ions with a general formulas [Mn(L)2(Cl)2]·4H2O (I), [Fe(L)2(Cl)2]·Cl·6H2O (II), [Cr(L)2(Cl)2]·Cl·6H2O (III) and [Zn(L)2(Cl)2]·2H2O (IV) where L = 4-acetylpyridine thiosemicarbazone, have been synthesized and interpreted using CHN elemental analysis, magnetic susceptibility measurements, molar conductance, thermal analysis and spectroscopic techniques; i.e., infrared, electronic UV/vis, 1H-NMR and mass. The manganese(II), ferric(III), chromium(III) and zinc(II) complexes have octahedral geometry. The molar conductance measurements reveal that the Mn(II) and Zn(II) chelates are non-electrolytes but Fe(III) and Cr(III) have an electrolytic behavior. The IR spectra show that the 4-acetylpyridine thiosemicarbazone free ligand is coordinated to the metal(II) chlorides as a neutral bidentate ligand through both of the lone pair of electrons of the C=N azomethine group and C=S group. X-ray powder diffraction gives an impression that the resulting complexes are amorphous and different from the start materials. The thermogravimetric studies indicate that uncoordinated water molecules are lost in the first and second decomposition steps. The activation thermodynamic parameters E*, ΔH*, ΔS* and ΔG* are estimated from the differential thermogravimetric analysis (DTG) curves using Horowitz–Metzger (HM) and Coats–Redfern (CR) methods. The ligand and its complexes have been screened for antibacterial and antifungal activities against two bacteria; i.e., Escherichia coli (Gram −ve) and Bacillus subtilis (Gram +ve) and two fungi, i.e., tricoderma and penicillium activities).  相似文献   

16.
Transparent monolithic gels were prepared from the cohydrolysis of diethylphosphatoethyltriethoxysilane (OC2H5)2P(O)CH2CH2Si(OC2H5)3 (SiP) and tetraethoxysilane Si(OC2H5)4 (TEOS) in various ratios, under acidic conditions. The hydrolysis and initial polycondensation of SiP were studied by high-resolution 29Si-NMR. Based on signal intensities, singnals due to the following oligomers were identified: R(OH)2SiOSi(OH)(R)OSi(OH)2, R(OH)2SiOSi(OH)2, R(OH)(OEt)SiOSi(OH)2R, R(OH)(OEt)SiOSi(OH)2R, R(OH)2SiOSi(OH)(R)OSi(OH)2R and [R(OH)SiO-]-3. 29Si NMR Quantitative analysis was performed on SiP–TEOS mixtures during the hydrolysis-condensation reactions and on the final gels. This provided information on the formation of the gels molecular structure. © 1996 John Wiley & Sons, Inc.  相似文献   

17.
Ning Kang  Jean-Christophe Leroux   《Polymer》2004,45(26):8967-8980
Novel A-B-A triblock and star-block amphiphilic copolymers, i.e. poly(N-(2-hydroxypropyl)methacrylamide)-block-poly(D,L-lactide)-block-poly(N-(2-hydroxypropyl)metha-crylamide), poly(N-vinyl-2-pyrrolidone)-block-poly(D,L-lactide)-block-poly(N-vinyl-2-pyrrolidone), star-poly(D,L-lactide)-block-poly(N-(2-hydroxypropyl)methacrylamide) and star-poly(D,L-lactide)-block-poly(N-vinylpyrrolidone), were synthesized and characterized. These polymers were prepared by free radical polymerization of N-(2-hydroxypropyl)methacrylamide and N-vinyl-2-pyrrolidone in the presence of either poly(D,L-lactide) dithiol or star-poly(D,L-lactide) tetrakis-thiol, both biodegradable macromolecular chain-transferring agents. All copolymers self-assembled in aqueous solution to form supramolecular aggregates of 20–180 nm in size. The critical aggregation concentration of the copolymers ranged from 5 to 24 mg/L, depending on their hydrophobicity. The partition equilibrium constant of pyrene in the hydrophobic core of micelles was between 0.71×105 and 1.63×105. The triblock copolymer micelles were loaded with two model poorly water-soluble drugs, namely, indomethacin (1.5–16.4% w/w) and paclitaxel (0.4–1.5% w/w), by a dialysis procedure. These triblock and star-block copolymers could prove useful as nanocarriers for the solubilization and delivery of hydrophobic drugs.  相似文献   

18.
Jeremy M. Rathfon 《Polymer》2008,49(7):1761-1769
Polymers exhibiting a thermoresponsive, lower critical solution temperature (LCST) phase transition have proven to be useful for many applications as “smart” or “intelligent” materials. A series of poly(N-isopropylmethacrylamide) (PNIPMAM) polymer, poly(N-isopropylmethacrylamide)-b-poly(acrylic acid) (PNIPMAM-b-PAA) diblock, and poly(acrylic acid)-b-poly(N-isopropylmethacrylamide)-b-poly(acrylic acid) (PAA-b-PNIPMAM-b-AA) triblock copolymer samples were synthesized via ATRP. A facile post-functionalization route was developed that uses an activated ester functionality to convert poly(N-methacryloxysuccinimide) (PMASI) blocks to LCST capable polyacrylamide, while poly(t-butyl acrylate) (PtBA) blocks were converted to water-soluble poly(acrylic acid) (PAA). The post-functionalization was monitored via 1H NMR and ATR-FTIR. The aqueous solution properties were explored and the PNIPMAM polymers were shown to have a LCST phase transition varying from 35 to 60 °C. The ability to synthesize block copolymers that are thermoresponsive and water-soluble will be of great benefit for broader applications in drug delivery, bioengineering, and nanotechnology.  相似文献   

19.
Two methods were used to remove Cr(VI) from industrial wastewater. Although both are based in the same general reaction: 3Fe(II)(aq) + Cr(VI)(aq) ; 3Fe(III)(aq) + Cr(III)(aq) the way in which the required amount of Fe(II) is added to the wastewater is different for each method. In the chemical method, Fe(II)(aq) is supplied by dissolving FeSO4 · 7(H2O)(s) into the wastewater, while in the electrochemical process Fe(II)(aq) ions are formed directly in solution by anodic dissolution of an steel electrode. After this reduction process, the resulting Cr(III)(aq) and Fe(III)(aq) ions are precipitated as insoluble hydroxide species, in both cases, changing the pH (i.e., adding Ca(OH)2(s)). Based on the chemical and thermodynamic characteristics of the systems Cr(VI)–Cr(III)–H2O–e and Fe(III)–Fe(II)–H2O–e both processes were optimized. However we show that the electrochemical option, apart from providing a better form of control, generates significantly less sludge as compared with the chemical process. Furthermore, it is also shown that sludge ageing promotes the formation of soluble polynuclear species of Cr(III). Therefore, it is recommended to separate the chromium and iron-bearing phases once they are formed. We propose the optimum hydraulic conditions for the continuous reduction of Cr(VI) present in the aqueous media treated in a plug-flow reactor.  相似文献   

20.
ABSTRACT

Solvent extraction of Eu(III) and Am(III) from weakly acidic solutions with octyl(phenyl)-N, N-diisobutylcarbamoylmethylphosphine oxide (CMPO) and 2,6-bis((diphenylphosphino)methyl)pyridine N, P, P-trioxide (NOPOPO) in 1,2-dichloroethane was studied on a comparative basis. NOPOPO was found to exhibit unusually high extractability for Eu(III) and Am(III), probably due to its trifunctional nature, sufficient steric flexibility and basicity of the functional groups. Both CMPO and NOPOPO demonstrated synergistic effects in extraction of Eu(III) and Am(III) when used in combination with thenoyltrifluoroacetone (HTTA). However, the stoichiometry of the extracted species with CMPO/HTTA and NOPOPO/HTTA was different under similar experimental conditions. The extractant dependencies of the synergistic extractions suggest that the extracted species are the adduct complexes, M(ClO4)(TTA)2(CMPO)2 and M(ClO4)2(TTA)(NOPOPO)2, respectively. It was also observed that CMPO and CMPO/HTTA in dichloroethane extracted Eu(III) and Am(III) equally well, with very similar extraction constants. However, NOPOPO and NOPOPO/HTTA in dichloroethane demonstrated a slight preference for Eu(III) over Am(III), with the extraction constants for Eu(III) more than one order of magnitude higher than that for Am(III).  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-23

京公网安备 11010802026262号