Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 18 de 18
Filter
1.
Phys Chem Chem Phys ; 16(47): 25770-4, 2014 Dec 21.
Article in English | MEDLINE | ID: mdl-25360625

ABSTRACT

We demonstrate the striking potential of exfoliated clay (synthetic hectorite; Laponite XLG) platelets to prepare bimetallic (Co-Pt and Ni-Pt) NCs with well-defined structures. Catalytic studies show a strong bimetallic synergistic effect of the core-shell NCs; their catalytic activities are much higher than those of monometallic NCs and other bimetallic core-shell NCs.

2.
Phys Chem Chem Phys ; 16(3): 880-4, 2014 Jan 21.
Article in English | MEDLINE | ID: mdl-24287459

ABSTRACT

We report a novel simple one-pot strategy for fabricating pure and highly planar silver (Ag) crystal belts. Unique single-crystal Ag belts (high width-to-thickness ratio ~50) were successfully synthesized in high yield (80 wt%) by reducing AgNO3 using N,N,N',N'-tetramethylethylenediamine (TEMED) as a reducing and a structure-determining agent in the presence of polyethylene glycol (PEG) under mild conditions.

3.
Langmuir ; 29(6): 1977-84, 2013 Feb 12.
Article in English | MEDLINE | ID: mdl-23343394

ABSTRACT

Novel and intriguing one-pot in situ method for the preparation of nanostructured Pt-clay materials under simple conditions is reported. In this synthesis, an inorganic clay mineral such as synthetic hectorite ("Laponite XLG") or natural montmorillonite ("Kunipia F") serves as a mild and effective reducing agent for Pt ions, which is uncommon for such a clay system, and also acts as an outstanding stabilizer for the resulting Pt nanoparticles. In aqueous solution, exfoliated colloidal clay platelets forms complex with Pt ions in the initial stage of mixing. Devoid of any organic dispersants or external reducing agents, subsequently, the Pt nanoparticles (3-6 nm) generated by clay-assisted in situ reduction of Pt ions successfully anchored onto the clay nanoplatelets. The Pt-clay material features a very high surface area (312 m(2) g(-1)) and has excellent catalytic activity, as was kinetically evaluated via the reduction of 4-nitrophenol with NaBH(4). After drying, this remarkably stable nanocomposite is completely redispersible in water and displays extreme thermal stability (up to 500 °C). On the basis of these results, this synthetic strategy is anticipated to be a very simple, economical, and green approach for the synthesis of nanostructured Pt-clay materials.

4.
Langmuir ; 29(10): 3214-22, 2013 Mar 12.
Article in English | MEDLINE | ID: mdl-23421650

ABSTRACT

We have investigated the adsorption and organization at the air/water interface of catanionic molecules released from a dispersion of solid-like catanionic vesicles composed of myristic acid and cetyl trimethylammonium chloride at the 2:1 ratio. These vesicles were shown recently to be promising foam stabilizers. Using Brewster angle microscopy, we observed the formation of a catanionic monolayer at the air/water interface composed of liquid-condensed domains in a liquid-expanded matrix. Further adsorption of catanionic molecules forced them to pack, thereby forming a very dense monolayer that prevented further vesicle rupture by avoiding contact of the vesicles with air. Moreover, confocal fluorescence microscopy revealed the presence of layers of intact vesicles that were progressively creaming toward this catanionic monolayer; the surface tension of the vesicle dispersion remained constant upon creaming. The catanionic monolayer behaved as a soft glassy material, an amorphous solid with time- and temperature-dependent properties. Using interfacial oscillatory rheology, we found that the monolayer relaxed mechanical stresses in seconds and melted at a temperature very close to the melting transition temperature of the vesicle bilayers. These results have potential application in the design of smart foams that have temperature-tunable stability.

5.
Phys Chem Chem Phys ; 15(39): 16477-80, 2013 Oct 21.
Article in English | MEDLINE | ID: mdl-24013787

ABSTRACT

Pt/fluorinated clay nanocomposites were prepared by clay-mediated in situ reduction; these nanocomposites demonstrated ultrahigh Pt nanoparticle thermal stability (up to 1000 °C), large BET surface area (379 m(2) g(-1)), uniform dispersion in aqueous media, and intriguing catalytic properties.

6.
Langmuir ; 26(5): 3115-20, 2010 Mar 02.
Article in English | MEDLINE | ID: mdl-19845315

ABSTRACT

Using small-angle X-ray scattering (SAXS), effects of different polyols used as polar additives, glycerol (GC), ethylene glycol (EG), and 1,2-butanediol (1,2-BD), on the structure of nonionic surfactant glycerol monoisostearate (iso-C(18)G(1)) reverse micelles in decane have been investigated as a function of polyol concentration and temperature. The real space structural functions of the reverse micelles were obtained by generalized indirect Fourier transformation (GIFT) evaluation of the SAXS data, letting the form factor virtually model-free, and the results were complemented by conventional theoretical model fittings. The iso-C(18)G(1) forms spheroid type or slightly elongated prolate type micelles in n-decane. We have found that addition of these polyols causes two-dimensional (2-D) growth of the reverse micelles in a similar manner to that induced by added water, both the maximum length and micellar cross-section diameter continuously increasing with polyol concentration. EG is most effective to increase the micellar size. 1,2-BD, which is apparently more hydrophobic than EG, has a weakest effect on the micellar growth. Unexpectedly, GC exhibits a less pronounced effect than EG despite its stronger hydrophilicity, which may be related to the similarity of the molecular structure to the surfactant glycerol moiety. The data demonstrate that different polarities of additives can be an additional tunable parameter for controlling the structure of the reverse micelles. The results well complement our recent findings that slightly different properties of solvent oils can provide a significant effect on the reverse micellar structures in glycerol-based surfactant systems.

7.
J Nanosci Nanotechnol ; 20(4): 2171-2178, 2020 04 01.
Article in English | MEDLINE | ID: mdl-31492226

ABSTRACT

We report a simple and spontaneous method for creation of extremely stable and catalytically active gold nanoparticles (AuNPs) from polyoxyethylene cholesteryl ether (ChEO15) and sodium tetrachloroaurate (III) dihydrate (Na[AuCl4] · 2H2O). AuNPs with uniform size and distribution were prepared at 27 °C (AuNPs were formed in <3 h) and at 50 °C (AuNPs were formed in <1 h). Our elucidations propose that creation of well-defined AuNPs is assisted by reductive ChEO15 surfactant solution via complexation and in-situ reduction of AuCl-4 ions in which ChEO15 acts both as a reducing and consequently a stabilizing (capping) moiety. AuNPs size can be altered by suitably choosing the Au precursor concentration. The resulting AuNPs reveal outstanding catalytic activity for conversion of nitrophenols to aminophenol in the presence of NaBH4 and also exhibit synthetic peroxidase-like activity. This simple yet effective synthesis technique opens prospect for creation of stable and active AuNPs.

8.
J Phys Chem B ; 111(35): 10438-47, 2007 Sep 06.
Article in English | MEDLINE | ID: mdl-17696527

ABSTRACT

We have studied the structure and rheological behavior of viscoelastic wormlike micellar solutions in the mixed nonionic surfactants poly(oxyethylene) cholesteryl ether (ChEO15)-trioxyethylene monododecyl ether (C12EO3) and anionic sodium dodecyl sulfate (SDS)-C12EO3 using a series of glycerol/water and formamide/water mixed solvents. The obtained results are compared with those reported in pure water for the corresponding mixed surfactant systems. The zero-shear viscosity first sharply increases with C12EO3 addition and then decreases; i.e., there is a viscosity maximum. The intensity (viscosity) and position (C12EO3 fraction) of this maximum shift to lower values upon an increase in the ratio of glycerol in the glycerol/water mixed solvent, while the position of the maximum changes in an opposite way with increasing formamide. In the case of the SDS/C12EO3 system, zero-shear viscosity shows a decrease with an increase of temperature, but for the ChEO15/C12EO3 system, again, the zero-shear viscosity shows a maximum if plotted as a function of temperature, its position depending on the C12EO3 mixing fraction. In the studied nonionic systems, worm micelles seem to exist at low temperatures (down to 0 degrees C) and high glycerol concentrations (up to 50 wt %), which is interesting from the viewpoint of applications such as drag reduction fluids. Rheology results are supported by small-angle X-ray scattering (SAXS) and dynamic light scattering (DLS) measurements on nonionic systems, which indicate micellar elongation upon addition of glycerol or increasing temperature and shortening upon addition of formamide. The results can be interpreted in terms of changes in the surface curvature of aggregates and lyophobicity.

9.
J Colloid Interface Sci ; 313(2): 680-5, 2007 Sep 15.
Article in English | MEDLINE | ID: mdl-17540398

ABSTRACT

A study of the phase and rheological behavior of anionic surfactant sodium dodecyl trioxyethylene sulfate (SDES) and nonionic polyoxyethylene sorbitan monooleate (Tween-80) with alkanoyl-N-methylethanolamide (C(12), NMEA-12; and C(16), NMEA-16) in aqueous system is presented. Upon addition of NMEA to the semi-dilute solution of SDES or Tween-80, induces micellar growth leading to the formation of a gel-like highly viscoelastic solution in the maximum viscosity region. These solutions obey the Maxwell model of a viscoelastic fluid. It was observed from rheological measurements that NMEA-16 is more effective than NMEA-12 to induce the micellar growth of surfactants. The relationship between the marked changes in viscosity with surfactant-cosurfactant mixing ratio based on the experimental observations is discussed.

10.
J Colloid Interface Sci ; 315(1): 330-6, 2007 Nov 01.
Article in English | MEDLINE | ID: mdl-17673245

ABSTRACT

The formation and rheological behavior of a viscoelastic wormlike micellar solution in an aqueous solution of a mixed surfactant system of alkyl ethoxylate sulfate (AES), C(12)H(25)(OCH(2)CH(2))(3)OSO(-)(3)Na(+), and polyoxyethylene dodecyl ether, C(12)EO(3), and the unusual effect of temperature on the rheological behavior have been studied. Upon successive addition of C(12)EO(3) to the dilute micellar solution of AES, viscosity increases swiftly and reaches its peak where a viscoelastic solution with nearly Maxwellian behavior is formed. With the further addition of C(12)EO(3), viscosity decreases sharply, which is attributed to the formation of micellar joints. With increasing temperature, the extent of micellar growth increases and the viscosity maximum is achieved at a lower mixing fraction of C(12)EO(3), but the maximum viscosity attained by the system decreases. The evolution of relaxation time and network density of the viscoelastic network also suggests that with increasing temperature, enhanced micellar growth takes place, but an additional, faster relaxation mechanism becomes increasingly favorable at high concentrations of C(12)EO(3). These results can be explained in terms of the increase in free energy of hemispherical end-caps (end-cap energy) of the micelles with increasing temperature.

11.
J Colloid Interface Sci ; 312(2): 489-97, 2007 Aug 15.
Article in English | MEDLINE | ID: mdl-17412356

ABSTRACT

Formation and rheological behaviour of viscoelastic wormlike micellar solution in aqueous mixed system of nonionic surfactants, polyoxyethylene sorbitan monooleate (Tween-80) and trioxyethylene alkyl ether (C(m)EO(3), m=12, 14, and 16) was studied. The semi-dilute aqueous solution of Tween-80, in presence of C(m)EO(3) shows a sharp increase in viscosity leading to the formation of a gel-like highly viscoelastic solution reaching a maximum, but decreases beyond an intermediate concentration and finally phase separates. When C(12)EO(3) is replaced by C(14)EO(3) the micellar growth occurs more effectively. However, with further increase in alkyl chain length, i.e. with C(16)EO(3), phase separation occurs before a viscoelastic solution is formed. The effect of temperature on the water/Tween-80/C(14)EO(3) mixed surfactant system was also studied. With increase in temperature, viscosity increases more promptly and the viscosity maximum is attained at relatively lower concentration of C(14)EO(3). Enhanced one dimensional micellar growth with increase in the temperature is mainly attributed to the decrease in the spontaneous curvature of the aggregates. Structural investigation by means of small-angle X-ray scattering (SAXS) technique of micelles in aqueous solution of Tween-80 and mixed water/Tween-80/C(m)EO(3) systems confirmed the one dimensional micellar growth in the above system.

12.
J Colloid Interface Sci ; 300(1): 338-47, 2006 Aug 01.
Article in English | MEDLINE | ID: mdl-16696999

ABSTRACT

This study provides new experimental evidence for the disconnection of the lamellar phase (L(alpha)) in pseudobinary water-non-ionic surfactant systems. To prove that the disconnection is indeed a general phenomenon the phase behavior of the pseudobinary system water-pentaethylene glycol dodecyl ether/hexaethylene glycol dodecyl ether (H(2)O-C(12)E(5)/C(12)E(6)) was investigated as a function of the surfactant composition delta and the total surfactant concentration gamma. At a fixed gamma of 0.10 the extension of the highly diluted L(alpha) phase shrank continuously with increasing amount of C(12)E(6), i.e., increasing delta, until it disappeared at delta=0.60. The gamma-T phase diagram of this particular surfactant mixture was found to have a disconnected L(alpha) phase. For the first time, SAXS measurements were carried out to monitor structural changes related to the disconnection. For this purpose the interlayer spacing d and the effective cross-sectional area a(s) were determined from the SAXS data along characteristic paths through the L(alpha) phase.

13.
Indian J Biochem Biophys ; 41(2-3): 107-12, 2004.
Article in English | MEDLINE | ID: mdl-22900338

ABSTRACT

Critical micelle concentrations (CMCs) of cationic surfactant (alkyltrimethylammonium bromides, CnTABr, where n = 10, 12, 14, 16 and 18), and a bile salt sodium cholate (NaC) were determined from surface tension, conductance and dye solubilization methods, while of their equimolar mixtures from surface tension and dye solubilization methods. The interaction parameter (beta) obtained from analysis of data, using Rubingh's theory showed strong interaction between NaC and cationic surfactant. Time-resolved fluorescence-quenching results revealed small-sized mixed spherical micelle with aggregation number much less than micelles of cationic surfactant.


Subject(s)
Cations/chemistry , Micelles , Sodium Cholate/chemistry , Surface-Active Agents/chemistry , Bile Acids and Salts/chemistry , Chemistry Techniques, Analytical , Chemistry, Pharmaceutical/methods , Coloring Agents/chemistry , Dose-Response Relationship, Drug , Models, Chemical , Surface Properties , Water/chemistry
14.
Chem Commun (Camb) ; 50(23): 3014-7, 2014 Mar 21.
Article in English | MEDLINE | ID: mdl-24457942

ABSTRACT

Clay (synthetic hectorite; Laponite XLG) plays a very crucial role in the formation and stabilization of core-shell nanocrystals and affords high stability, large BET surface area and stimulates the exceptional catalytic activity of the core-shell NCs.

15.
ACS Appl Mater Interfaces ; 5(22): 11613-7, 2013 Nov 27.
Article in English | MEDLINE | ID: mdl-24144319

ABSTRACT

Nanostructured novel Pt/Clay nanocomposites consisting of well-defined Pt nanoparticles prepared by clay-mediated in situ reduction displays very high thermal stability, large BET surface area and superior catalytic activity for CO oxidation as compared to a model reference Pt/SiO2 catalysts. CO oxidation has attracted renewed attention because of its technological importance in the area of pollution control. The Pt/Clay system consisting of Pt nanoparticles strongly immobilized between the atomic layers of clay inhibits nanoparticle sintering and loss of catalytic activity even after prolonged heating at high temperatures. At elevated temperatures (300 °C), the Pt/Clay system demonstrates significant enhancement of catalytic activity, with almost 100% CO conversion in less than 5 min. Emphasis is given to the role played by the clay supporting material which is chemically and thermally stable under the catalytic conditions of exhaust purification.

16.
Langmuir ; 25(8): 4435-42, 2009 Apr 21.
Article in English | MEDLINE | ID: mdl-19243155

ABSTRACT

Formation of reverse micelles in surfactant/oil binary systems without water addition and the tunable parameters for the structure control of such micelles are presented. The small-angle X-ray scattering (SAXS) technique has been used for the structural characterization of micelles. It has been found that the nonionic surfactant glycerol monoisostearate (abbreviated as iso-C18G1) forms reverse micelles in different organic solvents such as cyclohexane, n-decane, and n-hexadecane without the addition of water. The structure (shape and size) of the reverse micelles has been found to depend on the solvent nature (alkyl chain length of oil), composition, temperature, and added water. Phase behavior study has shown that iso-C18G1 forms isotropic single-phase solutions in the aforementioned oils at 25 degrees C. At lower temperatures (<20 degrees C) II phases (dispersion of solid or liquid crystal phase) has been observed. SAXS data were evaluated by the generalized indirect Fourier transformation (GIFT) method, which has drawn a clear picture on the structural variations of the reverse micellar aggregates. Small globular types of micelles are found in the iso-C18G1/cyclohexane system. On the other hand, elongated ellipsoidal prolatelike or rodlike micelles are found in iso-C18G1/decane or iso-C18G1/hexadecane systems. The underlying mechanism of this structural evolution may be explained in terms of the transfer free energy of hydrophilic glycerol moiety from hydrophilic to hydrophobic environment of oils with different chain lengths. Besides, the penetration of oils to the lipophilic chain of the surfactant in reverse micellar systems differs depending on the chain length of oils. Lowering temperature and increasing surfactant concentration similarly lead to micellar growth while the cross-section structure remains essentially unchanged. Addition of trace water induced micellar growth, which is accompanied by the rapid swelling of the micellar core. The results obtained by this study demonstrate that the solvent nature, temperature, composition, and water addition can be the tunable parameters for the size, shape, and internal structure control of the iso-C18G1-based reverse micelles.


Subject(s)
Glycerol/chemistry , Micelles , Oils/chemistry , Fourier Analysis , Models, Chemical , Models, Statistical , Scattering, Radiation , Solvents/chemistry , Surface Properties , Surface-Active Agents , Temperature , Water/chemistry , X-Rays
17.
J Colloid Interface Sci ; 325(1): 243-9, 2008 Sep 01.
Article in English | MEDLINE | ID: mdl-18632110

ABSTRACT

The solubilization of triglycerides [1,2,3-tributanoylglycerol (TBG) and 1,2,3-trihexanoylglycerol (THG)] in water/octa(oxyethylene) dodecyl ether (C(12)EO(8)) systems has been investigated. Oil-induced changes in the structure of liquid crystals in water/C(12)EO(8) system have been studied by optical observation and small-angle X-ray scattering (SAXS) measurements. In the water/C(12)EO(8)/oil systems, solubilization of THG and TBG induces a transition between H(1) (hexagonal) and L(alpha) (lamellar) liquid crystals at high C(12)EO(8) concentrations, whereas at low surfactant concentrations a H(1)-I(1) (discontinuous micellar cubic phase) transition occurs. This anomalous behavior is attributed to the partitioning of solubilized oil in the micelles. At low surfactant concentrations THG is mainly solubilized into the hydrophobic cores of the surfactant micelles, indicating high swelling or low penetration tendency, resulting in a steep increase in the radius of the aggregates (r(H)), thereby inducing a rod-sphere transition. At high surfactant concentrations, THG is not mainly solubilized into the core but distributed between the palisade layer and the core of the aggregates. The TBG is considerably solubilized into the surfactant palisade layer, indicating a high penetration tendency, resulting in an increase in the effective cross-sectional area per surfactant molecule, a(s). The thermal stability of the I(1) phase increases with the solubilization of THG into the aggregate cores. The percentage deviation of the experimental interlayer spacings (P(d)) from complete swelling was also evaluated for different triglycerides in the H(1) and L(alpha) phases or different surfactant concentrations. It is found that the penetration tendency of triglycerides could be used as a tuning parameter for I(1) phase formation depending on the surfactant concentration and the molecular weight of the oil.

18.
Langmuir ; 23(22): 11007-14, 2007 Oct 23.
Article in English | MEDLINE | ID: mdl-17910486

ABSTRACT

The structure of the reverse micellar cubic (I2) liquid crystal and the adjacent micellar phase in amphiphilic block copolymer/water/oil systems has been studied by small-angle X-ray scattering (SAXS), rheometry, and differential scanning calorimetry (DSC). Upon addition of water to the copolymer/oil mixture, spherical micelles are formed and grow in size until a disorder-order transition takes place, which is related to a sudden increase in the viscosity and shear modulus. The transition is driven by the packing of the spherical micelles into a Fd3m cubic lattice. The single-phase I2 liquid crystals show gel-like behavior and elastic moduli higher than 104 Pa, as determined by oscillatory measurements. Further addition of water induces phase separation, and it is found that reverse water-in-oil emulsions with high internal phase ratio and stabilized by I2 liquid crystals can be prepared in the two-phase region. Contrary to liquid-liquid emulsions, both the elastic modulus and the viscosity decrease with the fraction of dispersed water, due to a decrease in the crystalline fraction in the sample, although the reverse emulsions remain gel-like even at high volume fractions of the dispersed phase. A temperature induced order-disorder transition can be detected by calorimetry and rheometry. Upon heating the I2 liquid crystals, two thermal events associated with small enthalpy values were detected: one endothermic, related to the "melting" of the liquid crystal, and the other exothermic, attributed to phase separation. The melting of the liquid crystal is associated with a sudden drop in viscosity and shear moduli. Results are relevant for understanding the formation of cubic-phase-based reverse emulsions and for their application as templates for the synthesis of structured materials.

SELECTION OF CITATIONS
SEARCH DETAIL