Artículos de revistas sobre el tema "Micellar catalyst"

Siga este enlace para ver otros tipos de publicaciones sobre el tema: Micellar catalyst.

Crea una cita precisa en los estilos APA, MLA, Chicago, Harvard y otros

Elija tipo de fuente:

Consulte los 50 mejores artículos de revistas para su investigación sobre el tema "Micellar catalyst".

Junto a cada fuente en la lista de referencias hay un botón "Agregar a la bibliografía". Pulsa este botón, y generaremos automáticamente la referencia bibliográfica para la obra elegida en el estilo de cita que necesites: APA, MLA, Harvard, Vancouver, Chicago, etc.

También puede descargar el texto completo de la publicación académica en formato pdf y leer en línea su resumen siempre que esté disponible en los metadatos.

Explore artículos de revistas sobre una amplia variedad de disciplinas y organice su bibliografía correctamente.

1

Cibulka, Radek, Lenka Baxová, Hana Dvořáková, František Hampl, Petra Ménová, Viktor Mojr, Baptiste Plancq y Serkan Sayin. "Catalytic effect of alloxazinium and isoalloxazinium salts on oxidation of sulfides with hydrogen peroxide in micellar media". Collection of Czechoslovak Chemical Communications 74, n.º 6 (2009): 973–93. http://dx.doi.org/10.1135/cccc2009030.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Three novel amphiphilic alloxazinium salts were prepared: 3-dodecyl-5-ethyl-7,8,10-trimethylisoalloxazinium perchlorate (1c), 1-dodecyl-5-ethyl-3-methylalloxazinium perchlorate (2b), and 3-dodecyl-5-ethyl-1-methylalloxazinium perchlorate (2c). Their catalytic activity in thioanisole (3) oxidation with hydrogen peroxide was investigated in micelles of sodium dodecylsulfate (SDS), hexadecyltrimethylammonium nitrate (CTANO3) and Brij 35. Reaction rates were strongly dependent on the catalyst structure, on the type of micelles, and on pH value. Alloxazinium salts 2 were more effective catalysts than isoalloxazinium salts 1. Due to the contribution of micellar catalysis, the vcat/v0 ratio of the catalyzed and non-catalyzed reaction rates was almost 80 with salt 2b solubilized in CTANO3 micelles. Nevertheless, the highest acceleration was observed with non-amphiphilic 5-ethyl-1,3-dimethylalloxazinium perchlorate (2a) in CTANO3 micelles (vcat/v0 = 134). In this case, salt 2a presumably acts as a phase-transfer catalyst bringing hydrogen peroxide from the aqueous phase into the micelle interior. Synthetic applicability of the investigated catalytic systems was verified on semi-preparative scale.
2

Broxton, Trevor J. "Micellar Catalysis of Organic Reactions. XXXVIII A Study of the Catalytic Effect of Micelles of 3-Hydroxymethyl-1-tetradecylpyridinium Bromide on Amide Hydrolysis and Nucleophilic Aromatic Substitution". Australian Journal of Chemistry 51, n.º 7 (1998): 541. http://dx.doi.org/10.1071/c98053.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The preparation of 3-hydroxymethyl-1-tetradecylpyridinium bromide and its use as a catalyst of nucleophilic aromatic substitution and also amide hydrolysis are reported. It was found that the hydroxydehalogenation of nitro-activated aryl halides was much faster in these micelles than in the presence of cetyl(2-hydroxyethyl)dimethylammonium bromide. It was concluded that the increased catalysis of nucleophilic aromatic substitution by this micelle was due to a faster decomposition of the aryl micellar ether which must occur before the phenolate product is released. No such difference in the two micelles was found for amide or thioamide hydrolysis since in these reactions the product amine is produced in the first step of the reaction and decomposition of the acylated micelle is not required in the rate-determining step of the reaction.
3

Steven, Alan. "Micelle-Mediated Chemistry in Water for the Synthesis of Drug Candidates". Synthesis 51, n.º 13 (21 de mayo de 2019): 2632–47. http://dx.doi.org/10.1055/s-0037-1610714.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Micellar reaction conditions, in a predominantly aqueous medium, have been developed for transformations commonly used by synthetic chemists working in the pharmaceutical industry to discover and develop drug candidates. The reactions covered in this review are the Suzuki–Miyaura, Miyaura borylation, Sonogashira coupling, transition-metal-catalysed CAr–N coupling, SNAr, amidation, and nitro reduction. Pharmaceutically relevant examples of these applications will be used to show how micellar conditions can offer advantages in yield, operational ease, amount of waste generated, transition-metal catalyst loading, and safety over the use of organic solvents, irrespective of the setting in which they are used.1 Introduction2 Micelles as Solubilising Agents3 Micelles as Nanoreactors4 Designer Surfactants5 A Critical Evaluation of the Case for Chemistry in Micelles6 Scope of Review7 Suzuki–Miyaura Coupling8 Miyaura Borylation9 Sonogashira Coupling10 Transition-Metal-Catalysed CAr–N Couplings11 SNAr12 Amidation13 Nitro Reduction14 Micellar Sequences15 Summary and Outlook
4

Kuimov, Vladimir A., Svetlana F. Malysheva, Natalia A. Belogorlova, Ruslan I. Fattakhov, Alexander I. Albanov y Boris A. Trofimov. "Triton-X-100 as an Organic Catalyst for One-Pot Synthesis of Arylmethyl-H-phosphinic Acids from Red Phosphorus and Arylmethyl Halides in the KOH/H2O/Toluene Multiphase Superbase System". Catalysts 13, n.º 4 (11 de abril de 2023): 720. http://dx.doi.org/10.3390/catal13040720.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Triton-X-100, a polyethylene glycol 4-(tert-octyl)phenyl ether, has been found to be an active micellar organic catalyst for the one-pot selective synthesis of arylmethyl-H-phosphinic acids in up to 65% yields by the direct phosphinylation of arylmethyl halides with red phosphorus in the KOH/H2O/toluene multiphase superbase system. The catalyst demonstrates a good recyclability. As a result, an expeditious method for the chemoselective synthesis of arylmethyl-H-phosphinic acids—versatile sought-after organophosphorus compounds—has been developed. The synthesis is implemented via direct alkylation/oxidation of red phosphorus with arylmethyl halides, promoted by superbase hydroxide anions using Triton-X-100 (a commercial off-the-shelf organic recyclable micellar catalyst). The reaction comprises the hydroxide anions-assisted disassembly of Pred 3D polymer molecules triggered by the separation from the potassium cation in ordinary crown-like micelles to produce polyphosphide anions in aqueous phase. Further, polyphosphide anions are alkylated with arylmethyl halides in organic phase in the presence of the catalytic Triton-X-100 reverse micelles and alkylated polyphosphide species undergo the double hydroxylation. The advantages of the strategy developed include chemoselectivity, benign and accessible starting reagents, catalyst recyclability, and facile one-pot implementation.
5

Augustine, Rimesh, Dae-Kyoung Kim, Ho An Kim, Jae Ho Kim y Il Kim. "Poly(N-isopropylacrylamide)-b-Poly(L-lysine)-b-Poly(L-histidine) Triblock Amphiphilic Copolymer Nanomicelles for Dual-Responsive Anticancer Drug Delivery". Journal of Nanoscience and Nanotechnology 20, n.º 11 (1 de noviembre de 2020): 6959–67. http://dx.doi.org/10.1166/jnn.2020.18822.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
A series of ABC triblock poly(N-isopropylacrylamide)75-block-poly(L-lysine)35-block-poly(L-histidine)n (p(NIPAM)75-b-p(Lys)35-b-p(His)N) (N = 35,50,75,100) copolymer bio-conjugates were prepared by combining reversible addition-fragmentation chain transfer polymerization and fast ring-opening polymerization of N-carboxyanhydride a-amino acid using 1,3-dicyclohexylimidazolium hydrogen carbonate as a catalyst. All the resulting triblock copolymers self-assembled into spherical micellar aggregates in aqueous solution, irrespective of the chain length of the histidine block. The micellar aggregates encapsulated the anticancer drug doxorubicin (Dox) and exhibited high drug loading efficiency. Temperature and pH stimuli were applied to investigate the controlled release of Dox. The non-cytotoxic nature of the polymers was investigated using 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. Cellular uptake of the Dox-loaded micelles revealed that the micelles successfully release Dox in cancer cells in response to pH- and temperature-induced morphological change. In-vitro studies further confirmed that the Dox-loaded triblock copolymer micelle is an excellent platform for drug delivery.
6

Tang, Christina y Bridget T. McInnes. "Cascade Processes with Micellar Reaction Media: Recent Advances and Future Directions". Molecules 27, n.º 17 (31 de agosto de 2022): 5611. http://dx.doi.org/10.3390/molecules27175611.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Reducing the use of solvents is an important aim of green chemistry. Using micelles self-assembled from amphiphilic molecules dispersed in water (considered a green solvent) has facilitated reactions of organic compounds. When performing reactions in micelles, the hydrophobic effect can considerably accelerate apparent reaction rates, as well as enhance selectivity. Here, we review micellar reaction media and their potential role in sustainable chemical production. The focus of this review is applications of engineered amphiphilic systems for reactions (surface-active ionic liquids, designer surfactants, and block copolymers) as reaction media. Micelles are a versatile platform for performing a large array of organic chemistries using water as the bulk solvent. Building on this foundation, synthetic sequences combining several reaction steps in one pot have been developed. Telescoping multiple reactions can reduce solvent waste by limiting the volume of solvents, as well as eliminating purification processes. Thus, in particular, we review recent advances in “one-pot” multistep reactions achieved using micellar reaction media with potential applications in medicinal chemistry and agrochemistry. Photocatalyzed reactions in micellar reaction media are also discussed. In addition to the use of micelles, we emphasize the process (steps to isolate the product and reuse the catalyst).
7

Wood, Alex B., Daniel E. Roa, Fabrice Gallou y Bruce H. Lipshutz. "α-Arylation of (hetero)aryl ketones in aqueous surfactant media". Green Chemistry 23, n.º 13 (2021): 4858–65. http://dx.doi.org/10.1039/d1gc01572a.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
α-Arylations can be run under micellar catalysis conditions using a Pd(i) pre-catalyst together with KO-t-Bu as base. Sequences using this coupling along with as many as four additional steps can be carried out in a 1-pot fashion, all in water.
8

Razak, Norazizah Abd y M. Niyaz Khan. "Kinetics and Mechanism of Nanoparticles-Catalyzed Piperidinolysis of Anionic Phenyl Salicylate". Scientific World Journal 2014 (2014): 1–7. http://dx.doi.org/10.1155/2014/604139.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The values of the relative counterion (X) binding constantRXBr(=KX/KBr, whereKXandKBrrepresent cetyltrimethylammonium bromide, CTABr, micellar binding constants ofXv-(in non-spherical micelles),v=1,2, and Br−(in spherical micelles)) are 58, 68, 127, and 125 forXv−=1−, 12−, 2−, and22-, respectively. The values of 15 mM CTABr/[NavX] nanoparticles-catalyzed apparent second-order rate constants for piperidinolysis of ionized phenyl salicylate at 35°C are 0.417, 0.488, 0.926, and 0.891 M−1 s−1forNavX= Na1, Na21, Na2, and Na22, respectively. Almost entire catalytic effect of nanoparticles catalyst is due to the ability of nonreactive counterions,Xv-, to expel reactive counterions,3−, from nanoparticles to the bulk water phase.
9

Schwarze, M., M. Schmidt, L. A. T. Nguyen, A. Drews, M. Kraume y R. Schomäcker. "Micellar enhanced ultrafiltration of a rhodium catalyst". Journal of Membrane Science 421-422 (diciembre de 2012): 165–71. http://dx.doi.org/10.1016/j.memsci.2012.07.017.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
10

LINDKVIST, Björn, Rolf WEINANDER, Lars ENGMAN, Marc KOETSE, Jan B. F. N. ENGBERTS y Ralf MORGENSTERN. "Glutathione transferase mimics: micellar catalysis of an enzymic reaction". Biochemical Journal 323, n.º 1 (1 de abril de 1997): 39–43. http://dx.doi.org/10.1042/bj3230039.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Substances that mimic the enzyme action of glutathione transferases (which serve in detoxification) are described. These micellar catalysts enhance the reaction rate between thiols and activated halogenated nitroarenes as well as α,β-unsaturated carbonyls. The nucleophilic aromatic substitution reaction is enhanced by the following surfactants in descending order: poly(dimethyldiallylammonium-co-dodecylmethyldiallylammon ium) bromide (86/14) ≫cetyltrimethylammonium bromide > zwittergent 3-16 (n-hexadecyl-N,N-dimethyl-3-ammonio-1-propanesulphonate) > zwittergent 3-14 (n-tetradecyl-N,N-dimethyl-3-ammonio-1-propanesulphonate) ≈ N,N-dimethyl-laurylamine N-oxide > N,N-dimethyloctylamine N-oxide. The most efficient catalyst studied is a polymeric material that incorporates surfactant properties (n-dodecylmethyldiallylammonium bromide) and opens up possibilities for engineering sequences of reactions on a polymeric support. Michael addition to α,β-unsaturated carbonyls is exemplified by a model substance, trans-4-phenylbut-3-en-2-one, and a toxic compound that is formed during oxidative stress, 4-hydroxy-2-undecenal. The latter compound is conjugated with the highest efficiency of those tested. Micellar catalysts can thus be viewed as simple models for the glutathione transferases highlighting the influence of a positive electrostatic field and a non-specific hydrophobic binding site, pertaining to two catalytic aspects, namely thiolate anion stabilization and solvent shielding.
11

Jin, Bo, Fabrice Gallou, John Reilly y Bruce H. Lipshutz. "ppm Pd-catalyzed, Cu-free Sonogashira couplings in water using commercially available catalyst precursors". Chemical Science 10, n.º 12 (2019): 3481–85. http://dx.doi.org/10.1039/c8sc05618h.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
A new catalyst that derives from commercially available precursors for copper-free, Pd-catalyzed Sonogashira reactions at the sustainable ppm level of precious metal palladium under mild aqueous micellar conditions has been developed.
12

Khan, Mohammad Niyaz y Ibrahim Isah Fagge. "Kinetics and Mechanism of Cationic Micelle/Flexible Nanoparticle Catalysis: A Review". Progress in Reaction Kinetics and Mechanism 43, n.º 1 (marzo de 2018): 1–20. http://dx.doi.org/10.3184/146867818x15066862094905.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The aqueous surfactant (Surf) solution at [Surf] > cmc (critical micelle concentration) contains flexible micelles/nanoparticles. These particles form a pseudophase of different shapes and sizes where the medium polarity decreases as the distance increases from the exterior region of the interface of the Surf/H2O particle towards its furthest interior region. Flexible nanoparticles (FNs) catalyse a variety of chemical and biochemical reactions. FN catalysis involves both positive catalysis ( i.e. rate increase) and negative catalysis ( i.e. rate decrease). This article describes the mechanistic details of these catalyses at the molecular level, which reveals the molecular origin of these catalyses. Effects of inert counterionic salts (MX) on the rates of bimolecular reactions (with one of the reactants as reactive counterion) in the presence of ionic FNs/micelles may result in either positive or negative catalysis. The kinetics of cationic FN (Surf/MX/H2O)-catalysed bimolecular reactions (with nonionic and anionic reactants) provide kinetic parameters which can be used to determine an ion exchange constant or the ratio of the binding constants of counterions.
13

Tasca, Elena, Giorgio La Sorella, Laura Sperni, Giorgio Strukul y Alessandro Scarso. "Micellar promoted multi-component synthesis of 1,2,3-triazoles in water at room temperature". Green Chemistry 17, n.º 3 (2015): 1414–22. http://dx.doi.org/10.1039/c4gc02248c.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Micellar media in water provide a simple and efficient environment favoring the multi-component synthesis of 1,2,3-triazoles from organic bromides, sodium azide and terminal alkynes in the presence of [Cu(IMes)Cl] 1 catalyst at room temperature within a few hours.
14

Belousova, I. A., T. M. Prokopyeva, N. G. Razumova, T. S. Gaidash y V. A. Mikhailov. "Turnover in acyl substrates destruction in organized microheterogeneous systems". Vestnik NovSU, n.º 3 (2023): 346–56. http://dx.doi.org/10.34680/2076-8052.2023.3(132).346-356.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Design of organized microheterogeneous systems for fast disrupting (disintegration) of acylcontaining substrates (including organophosphorus compounds) is based on the use of highly reactive compounds. Reagents (catalysts) ability to provide “turnover” is of great impotence. Kinetic regularities for 4- nitrophenylacetate (PNPA) deacylation with 1-methyl-3-(2-hydroximinoethyl)-imidazolium chloride (oxime), chloral, and their mixtures were studied in water and cetyltrimethylammonium bromide (CTAB) solutions, under pH = const and [CTAB] = 10-2 M. Rate constants dependences upon nucleophile (catalyst) and detergent concentration are typical for reactions in water and micellar pseudophase. Transferring of PNPA deacylation from water into surfactant micelles leads to three orders rate enhancement. Kinetic experiments at different substrate concentrations indicate that a) first order rate constant in the chloral-CTAB system does not change up to tenfold excess of PNPA; b) first order rate constant in the oxime-CTAB system decreases for approx. 40% at PNPA excess; c) all advantages of “turnover” are achievable in the system chloral-oximeCTAB, where chloral provides first disintegration of acylated oxime with generation of highly reactive oximate-ion. Results obtained pave the way to modifying organized microheterogeneous systems, first of all, by surfactant structure changes.
15

Takale, Balaram S., Ruchita R. Thakore, Sachin Handa, Fabrice Gallou, John Reilly y Bruce H. Lipshutz. "A new, substituted palladacycle for ppm level Pd-catalyzed Suzuki–Miyaura cross couplings in water". Chemical Science 10, n.º 38 (2019): 8825–31. http://dx.doi.org/10.1039/c9sc02528f.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
A newly engineered palladacycle that contains substituents on the biphenyl rings along with the ligand HandaPhos is especially well-matched to an aqueous micellar medium, enabling valued Suzuki–Miyaura couplings to be run not only in water under mild conditions, but at 300 ppm of Pd catalyst.
16

Akporji, Nnamdi, Ruchita R. Thakore, Margery Cortes-Clerget, Joel Andersen, Evan Landstrom, Donald H. Aue, Fabrice Gallou y Bruce H. Lipshutz. "N2Phos – an easily made, highly effective ligand designed for ppm level Pd-catalyzed Suzuki–Miyaura cross couplings in water". Chemical Science 11, n.º 20 (2020): 5205–12. http://dx.doi.org/10.1039/d0sc00968g.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
A new, biaryl phosphine-containing ligand, N2Phos, forms a 1 : 1 complex with Pd resulting in an active catalyst at the ppm level for Suzuki–Miyaura couplings in water, enabled by an aqueous micellar medium. Notably, aryl chlorides are shown to be amenable substrates.
17

Pasricha, Sharda. "Aqueous Phase Bromination by Micellar Solution of Sodium Dodecyl Sulfate (SDS): An Undergraduate Chemistry Experiment". Current Catalysis 10, n.º 3 (diciembre de 2021): 214–18. http://dx.doi.org/10.2174/2211544710666211119100631.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Background: Bromination is a key reaction in the chemical industry since the organobromines find application in diverse fields like pharmaceuticals, dyes, fire retardants, and as intermediates in chemical synthesis. Objective: To carry out green, in-situ bromination of acetanilide in an aqueous medium using micellar SDS as a catalyst. Methods: Bromination of acetanilide in-situ using potassium bromide as a non-corrosive source of bromine, ceric ammonium nitrate as oxidant, micellar solution of sodium dodecyl sulphate (SDS) as catalyst and water as solvent. Results: p-Bromoacetanilide was prepared in excellent yields at room temperature using green chemistry principles. Conclusion: The presented method provides a fast and environmentally safe route for the preparation of p-bromoacetanilide from acetanilide. It avoids the use of volatile, corrosive, and hazardous substances like liquid bromine and acetic acid. The use of water makes it safer and free from hazardous organic solvents. This reaction can be suitably adopted at the undergraduate level and may find use in the synthesis of commercially important bromo compounds.
18

Ahanthem, Dini, Devi Prasan Ojha, Francis A. S. Chipem y Warjeet S. Laitonjam. "One-pot Pseudo-Domino Three-Component Knoevenagel Condensation Reaction in Water Enabled by Micellar Catalyst: Mechanism and Reactivity". Letters in Organic Chemistry 17, n.º 11 (29 de noviembre de 2020): 823–31. http://dx.doi.org/10.2174/1570178616666190701102542.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The micellar catalysis is well-known for its hydrophobic effect that is the tendency of nonpolar groups to cluster within the lipophilic core so as to shield them from contact with an aqueous environment formed upon the dissolution of a surfactant in water. This provides a unique opportunity to establish organic transformations in greener solvents, such as water leading to organic waste control and easy product isolation protocols. Considering the significant interaction of thiobarbituric acid moieties in a biological macromolecule core, herein, a highly efficient procedure for the synthesis of biological and medicinal important 5-(arylmethylene)dihydro-2-thioxo-4,6(1H,5H)-pyrimidinediones via Knoevenagel condensation of thiobarbituric acids and aldehydes catalyzed by a surfactant, sodium dodecyl sulfate, is developed. The synthetic procedure shows the excellent activity of the micellar catalysts towards the aldehyde activation leading to a facile condensation. The application of the method is demonstrated by further synthesis of 5,5'-(4-arylmethylene)bis[dihydro-2-thioxo-4,6(1H,5H)- pyrimidinediones]. Theoretical studies of the reaction were also carried out to investigate the effect of electron releasing and electron-withdrawing group in benzaldehyde on the reaction.
19

Schwarze, Michael, Anke Rost, Thomas Weigel y Reinhard Schomäcker. "Selection of systems for catalyst recovery by micellar enhanced ultrafiltration". Chemical Engineering and Processing: Process Intensification 48, n.º 1 (enero de 2009): 356–63. http://dx.doi.org/10.1016/j.cep.2008.04.014.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
20

Liang, Chunshuang y Shimei Jiang. "Fluorescence light-up detection of cyanide in water based on cyclization reaction followed by ESIPT and AIEE". Analyst 142, n.º 24 (2017): 4825–33. http://dx.doi.org/10.1039/c7an01479a.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Sensor 1 could undergo an oxidative cyclization reaction and then generate hydroxyphenylbenzoxazole 2 when CN was present as a catalyst. The cyclization product 2 displayed fluorescence properties in the micellar due to the AIEE as well as ESIPT effect. This reaction process could be used for the light-up detection of CN in water.
21

Yusuf, Osman, Raisuddin Ali, Abdullah H. Alomrani, Aws Alshamsan, Abdullah K. Alshememry, Abdulaziz M. Almalik, Afsaneh Lavasanifar y Ziyad Binkhathlan. "Design and Development of D‒α‒Tocopheryl Polyethylene Glycol Succinate‒block‒Poly(ε-Caprolactone) (TPGS−b−PCL) Nanocarriers for Solubilization and Controlled Release of Paclitaxel". Molecules 26, n.º 9 (4 de mayo de 2021): 2690. http://dx.doi.org/10.3390/molecules26092690.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The objective of this study was to synthesize and characterize a set of biodegradable block copolymers based on TPGS-block-poly(ε-caprolactone) (TPGS-b-PCL) and to assess their self-assembled structures as a nanodelivery system for paclitaxel (PAX). The conjugation of PCL to TPGS was hypothesized to increase the stability and the drug solubilization characteristics of TPGS micelles. TPGS-b-PCL copolymer with various PCL/TPGS ratios were synthesized via ring opening bulk polymerization of ε-caprolactone using TPGS, with different molecular weights of PEG (1–5 kDa), as initiators and stannous octoate as a catalyst. The synthesized copolymers were characterized using 1H NMR, GPC, FTIR, XRD, and DSC. Assembly of block copolymers was achieved via the cosolvent evaporation method. The self-assembled structures were characterized for their size, polydispersity, and CMC using dynamic light scattering (DLS) technique. The results from the spectroscopic and thermal analyses confirmed the successful synthesis of the copolymers. Only copolymers that consisted of TPGS with PEG molecular weights ≥ 2000 Da were able to self-assemble and form nanocarriers of ≤200 nm in diameter. Moreover, TPGS2000-b-PCL4000, TPGS3500-b-PCL7000, and TPGS5000-b-PCL15000 micelles enhanced the aqueous solubility of PAX from 0.3 µg/mL up to 88.4 ug/mL in TPGS5000-b-PCL15000. Of the abovementioned micellar formulations, TPGS5000-b-PCL15000 showed the slowest in vitro release of PAX. Specifically, the PAX-loaded TPGS5000-b-PCL15000 micellar formulation showed less than 10% drug release within the first 12 h, and around 36% cumulative drug release within 72 h compared to 61% and 100% PAX release, respectively, from the commercially available formulation (Ebetaxel®) at the same time points. Our results point to a great potential for TPGS-b-PCL micelles to efficiently solubilize and control the release of PAX.
22

Schmidt, Fabian, Bastian Zehner, Marlene Kaposi, Markus Drees, János Mink, Wolfgang Korth, Andreas Jess y Mirza Cokoja. "Activation of hydrogen peroxide by the nitrate anion in micellar media". Green Chemistry 23, n.º 5 (2021): 1965–71. http://dx.doi.org/10.1039/d0gc03497e.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Surface-active imidazolium nitrates activate hydrogen peroxide, which enables the epoxidation of olefins. The micelles solubilise the substrate in the aqueous oxidant phase and allow for simple product separation and catalyst recycling.
23

Drennan, Catherine E., Rachelle J. Hughes, Vincent C. Reinsborough y Oladega O. Soriyan. "Article". Canadian Journal of Chemistry 76, n.º 2 (1 de febrero de 1998): 152–57. http://dx.doi.org/10.1139/v97-226.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Kinetic studies through stopped-flow spectroscopy were undertaken in the dilute solution range of anionic surfactants where pronounced rate enhancement or inhibition of Ni2+-ligand complexations is often observed at surfactant concentrations much below the critical micelle concentration (CMC). The results are interpreted in terms of Ni-surfactant micelles as the agents responsible for the rate changes in dilute surfactant solution. At higher surfactant concentrations these micelles are transformed into mixed micelles (counterion and size changes), eventually becoming normal surfactant micelles close to the CMC. Surface tension, dye solubility, conductivity, and fluorescent probe investigations support this interpretation.Key words: micellar catalysis, sodium dodecyl sulfate, micelles, critical micelle concentration, premicelles, Ni2+-ligand complexations.
24

Kulič, Jiří y Aleš Ptáček. "Alkaline Hydrolysis of 4-Nitrophenyl Acetate and Diphenyl (4-Nitrophenyl) Phosphate Catalyzed by Iodosoarenesulfonic Acids". Collection of Czechoslovak Chemical Communications 59, n.º 6 (1994): 1392–99. http://dx.doi.org/10.1135/cccc19941392.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Alkaline hydrolysis of 4-nitrophenyl acetate and of diphenyl (4-nitrophenyl) phosphate catalyzed by 2-iodosobenzenesulfonic and 2-iodoso-1-naphthalenesulfonic acids was studied in the presence of hexadecyltrimethylammonium bromide as the micellar agent. It was found that 2-iodosobenzenesulfonic acid is the better catalyst for the hydrolysis of phenyl acetate while 2-iodoso-1-naphthalenesulfonic acid is more efficient for the hydrolysis of the phosphate.
25

Houyi, N., S. Taichenc y L. Ganzuo. "PREPARATION OF MONODISPERSE NICKEL (COBALT) BOR1DE CATALYST USING REVERSED MICELLAR SYSTEM". Journal of Dispersion Science and Technology 13, n.º 6 (diciembre de 1992): 647–56. http://dx.doi.org/10.1080/01932699208943344.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
26

Király, Zoltán, Bernadett Veisz, Imre Dékány, Ágnes Mastalir y Zsolt Rázga. "Preparation of an organophilic palladium montmorillonite catalyst in a micellar system". Chemical Communications, n.º 19 (1999): 1925–26. http://dx.doi.org/10.1039/a905321b.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
27

Yousif, Dawod, Silvia Tombolato, Elmehdi Ould Maina, Riccardo Po, Paolo Biagini, Antonio Papagni y Luca Vaghi. "Micellar Suzuki Cross-Coupling between Thiophene and Aniline in Water and under Air". Organics 2, n.º 4 (16 de diciembre de 2021): 415–23. http://dx.doi.org/10.3390/org2040025.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The Suzuki–Miyaura cross-coupling reaction plays a fundamental role in modern synthetic organic chemistry, both in academia and industry. For this reason, scientists continue to search for new, more effective, cheaper and environmentally friendly procedures. Recently, micellar synthetic chemistry has been demonstrated to be an excellent strategy for achieving chemical transformations in a more efficient way, thanks to the creation of nanoreactors in aqueous environments using selected surfactants. In particular, the cheap and commercially available surfactant Kolliphor EL (a polyethoxylated castor oil derivative) has been used with success to achieve metal-catalyzed transformations in water with high yields and short reaction times, with the advantage of using air-sensitive catalysts without the need for inert atmosphere. In this work, the Kolliphor EL methodology was applied to the Suzuki cross-coupling reaction between thiophene and aniline, using the highly effective catalyst Pd(dtbpf)Cl2. The cross-coupling products were achieved at up to 98% yield, with reaction times of up to only 15 min, working at room temperature and without the need for inert atmosphere.
28

Yu, Xiaoqian, Artjom Herberg y Dirk Kuckling. "Micellar Organocatalysis Using Smart Polymer Supports: Influence of Thermoresponsive Self-Assembly on Catalytic Activity". Polymers 12, n.º 10 (1 de octubre de 2020): 2265. http://dx.doi.org/10.3390/polym12102265.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Micellar catalysts with a switchable core are attractive materials in organic synthesis. However, little is known about the role of the shell forming block on the performance of the catalyst. Thermoresponsive block copolymers based on poly(N-isopropylacrylamide-co-vinyl-4,4-dimethylazlactone) attached to different permanently hydrophilic blocks, namely poly(ethylene glycol), poly(N,N-dimethylacrylamide), and poly(2,3-dihydroxypropyl acrylate), were successfully synthesized via reversible addition/fragmentation chain transfer radical polymerization (RAFT). Post-polymerization attachment of an amino-functionalized L-prolineamide using the azlactone ring-opening reaction afforded functionalized thermoresponsive block copolymers. Temperature-induced aggregation of the functionalized block copolymers was studied using dynamic light scattering. It was shown that the chemical structure of the permanently hydrophilic block significantly affected the size of the polymer self-assemblies. The functionalized block copolymers were subjected to an aldol reaction between p-nitrobenzaldehyde and cyclohexanone in water. Upon temperature-induced aggregation, an increase in conversion was observed. The enantioselectivity of the polymer-bound organocatalyst improved with an increasing hydrophilic/hydrophobic interface as a result of the different stability of the polymer aggregates.
29

Dahadha, Adnan A., Mohammed Hassan, Tamara Mfarej, Razan Bani Issa, Mohamed J. Saadh, Mohammad Al-Dhoun, Mohammad Abunuwar y Nesrin T. Talat. "The Catalytic Influence of Polymers and Surfactants on the Rate Constants of Reaction of Maltose with Cerium (IV) in Acidic Aqueous Medium". Journal of Chemistry 2022 (1 de julio de 2022): 1–11. http://dx.doi.org/10.1155/2022/2609478.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Kinetics of the reaction of maltose with cerium ammonium sulfate were analyzed spectrophotometrically by observing the decrease of the absorbance of cerium (IV) at 385 nm in the presence and absence of polyethylene glycols (600, 1500, and 4000) and polyvinylpyrrolidone (PVP), in addition to anionic micelles of sodium dodecyl sulfate (SDS), cationic micelles of cetyltrimethylammonium bromide (CTAB) and non-ionic micelles of Tween 20 surfactants. Generally, there is little literature about using the polymers (PEGs and PVP) as catalysts in the oxidation-reduction reactions. Therefore, the major target of this work was to investigate the influence of the nature of polymers and surfactants on the oxidation rates of maltose by cerium (IV) in acidic aqueous media, as well as employing the Piszkiewicz model to explain the catalytic effect. The kinetic runs were derived by adaptation of the pseudo first-order reaction conditions with respect to the cerium (IV). The reaction was found to be first-order with respect to the oxidant and fractional-order to maltose and H2SO4. The reaction rates were enhanced in the presence of polymer and micellar catalysis. Indeed, the surfactants were found to work perfectly close to their critical micelle concentrations (CMC). Electrostatic interaction and H-bonding appear to play an influential role in binding maltose molecules to polymer/surfactant micelles, while oxidant ions remain at the periphery of the Stern layer within the micelle.
30

Broxton, TJ, JR Christie y RPT Chung. "Micellar Catalysis of Organic Reactions. XXVI. SNAr Reactions of Azide Ions". Australian Journal of Chemistry 42, n.º 6 (1989): 855. http://dx.doi.org/10.1071/ch9890855.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The azidodehalogenation of a number of aromatic compounds has been studied in the presence of micelles of cetyltrimethylammonium bromide (ctab). The variation of the observed rate of reaction with ctab concentration has been treated by using the model of Rodenas and Vera to determine the rate constant for reaction in the micellar pseudo-phase, k2m, the binding constant of the substrate to the micelle, Ks, and the nucleophile-micellar counter ion exchange constant KAzBr :. The ratio of the rate constants in the micellar pseudo-phase and in water varied between 0.9 and 52. For reactions involving the production of a dianionic intermediate the largest catalysis was found for compounds containing two nitro groups to stabilize the double negative charge. In addition significant differences in the catalysis were found between compounds having the reaction centre at the micelle-water interface and those for which the reaction centre was more buried inside the micelle. As previously reported the resulting aryl azides undergo cyclization to form a benzofuroxan if a nitro group is located ortho to the azide group. Furthermore, a reversible photochemical reaction was detected for two compounds having a carboxylate group ortho to the azide group.
31

Kulič, Jiří y Aleš Ptáček. "Catalyzed Alkaline Hydrolysis of Substituted Phenyl Acetates". Collection of Czechoslovak Chemical Communications 58, n.º 8 (1993): 1798–802. http://dx.doi.org/10.1135/cccc19931798.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
2-Iodosobenzoic acid forming strong nucleophile in alkaline medium - 1-oxido-1,2-benziodoxol-3(1H)-one, was used as a catalyst of alkaline hydrolysis of substituted phenyl acetates (4-NO2, 3-NO2, 3-Cl, 4-Br, H, 4-CH3, 3-CH3, 4-OCH3, 3-OCH3) in the presence of hexadecyltrimethylammonium bromide as a micellar agent. It was found that the observed first-order rate constants kobs can be correlated by the Hammett equation: log kobs = (-3.29 ± 0.03) + (1.77 ± 0.001) σ.
32

Gebicka, Lidia y Monika Jurgas-Grudzinska. "Activity and Stability of Catalase in Nonionic Micellar and Reverse Micellar Systems". Zeitschrift für Naturforschung C 59, n.º 11-12 (1 de diciembre de 2004): 887–91. http://dx.doi.org/10.1515/znc-2004-11-1220.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Catalase activity and stability in the presence of simple micelles of Brij 35 and entrapped in reverse micelles of Brij 30 have been studied. The enzyme retains full activity in aqueous micellar solution of Brij 35. Catalase exhibits “superactivity” in reverse micelles composed of 0.1 ᴍ Brij 30 in dodecane, n-heptane or isooctane, and significantly lowers the activity in decaline. The incorporation of catalase into Brij 30 reverse micelles enhances its stability at 50 °C. However, the stability of catalase incubated at 37 °C in micellar and reverse micellar solutions is lower than that in homogeneous aqueous solution.
33

Oranli, Levent, Pratap Bahadur y Gérard Riess. "Hydrodynamic studies on micellar solutions of styrene–butadiene block copolymers in selective solvents". Canadian Journal of Chemistry 63, n.º 10 (1 de octubre de 1985): 2691–96. http://dx.doi.org/10.1139/v85-447.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Hydrodynamic radius of micelles of several block copolymers in different selective solvents (for both types of blocks) was determined from photon correlation spectroscopy. The boundaries of micellar solutions in heptane (good solvent for polybutadiene block) and dimethylformamide (good solvent for polystyrene block) were established for polymers in terms of their molecular mass and block composition. The photon correlation spectroscopy data in combination with intrinsic viscosities of block copolymers in selective solvents were used to determine micellar molecular mass and aggregation number. The influence of temperature on the micelle size was examined. The block copolymer micelles could solubilize a certain amount of insoluble homopolymer within their insoluble core. 1H nmr spectra were examined to study the influence of temperature on micellar systems.
34

Wasylishen, Roderick E., Jan C. T. Kwak, Zhisheng Gao, Elisabeth Verpoorte, J. Bruce MacDonald y Ross M. Dickson. "NMR studies of hydrocarbons solubilized in aqueous micellar solutions". Canadian Journal of Chemistry 69, n.º 5 (1 de mayo de 1991): 822–33. http://dx.doi.org/10.1139/v91-122.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Information concerning the solubilization of hydrocarbons in ionic surfactant micelles was obtained from 2H NMR relaxation, 1H NMR chemical shifts, and 1H NMR paramagnetic relaxation measurements. The rotational motion of deuterated hydrocarbons, which is related to the micellar microviscosity at the location of the hydrocarbons, was probed by 2H NMR relaxation. The relaxation data are interpreted using both the two-step and the single-step models, and the results are discussed in terms of the micellar microviscosity and the location of the hydrocarbons in micelles. The location of the hydrocarbons in micelles was further investigated by determining the aromatic ring current-induced 1H chemical shifts along the surfactant alkyl chain and by comparing the 1H spin-lattice relaxation enhancement of the hydrocarbons and the surfactant alkyl chain, induced by Mn2+ on the micellar surface. The hydrocarbons used include benzene, naphthalene, acenaphthalene, triphenylene, cyclohexane, cyclododecane, and tert-butylcyclohexane and the surfactants studied are hexadecyl-, tetradecyl-, and dodecyltrimethylammonium bromide; hexadecyl-, tetradecyl-, and dodecylpyridinium halide; and sodium dodecyl sulfate. The results indicate that the micellar microviscosity at the location of saturated hydrocarbons is approximately 5 cP for both the cationic and anionic micelles, whereas the micellar microviscosity at the location of unsaturated hydrocarbons is much higher. The unsaturated hydrocarbons are found to reside primarily near the surfactant headgroup in the cationic micelles, but are distributed evenly throughout the anionic SDS micelles. The saturated hydrocarbons appear to be located in the interior of the micelles. Key words: NMR, relaxation, solubilization, surfactant, micelle.
35

Zehner, Bastian, Wolfgang Korth, Fabian Schmidt, Mirza Cokoja y Andreas Jess. "Kinetics of Epoxidation of Cyclooctene with Ionic Liquids Containing Tungstate as Micellar Catalyst". Chemical Engineering & Technology 44, n.º 12 (5 de noviembre de 2021): 2374–81. http://dx.doi.org/10.1002/ceat.202100102.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
36

Huang, Xin, Zeyuan Dong, Junqiu Liu, Shizhong Mao, Jiayun Xu, Guimin Luo y Jiacong Shen. "Selenium-Mediated Micellar Catalyst: An Efficient Enzyme Model for Glutathione Peroxidase-like Catalysis". Langmuir 23, n.º 3 (enero de 2007): 1518–22. http://dx.doi.org/10.1021/la061727p.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
37

Schmidt, Marcel, Saskia Schreiber, Luise Franz, Hauke Langhoff, Ashkan Farhang, Moritz Horstmann, Hans-Joachim Drexler, Detlef Heller y Michael Schwarze. "Hydrogenation of Itaconic Acid in Micellar Solutions: Catalyst Recycling with Cloud Point Extraction?" Industrial & Engineering Chemistry Research 58, n.º 7 (17 de septiembre de 2018): 2445–53. http://dx.doi.org/10.1021/acs.iecr.8b03313.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
38

Yin, Hong, Qiushi Wang, Sebastian Geburt, Steffen Milz, Bart Ruttens, Giedrius Degutis, Jan D'Haen et al. "Controlled synthesis of ultrathin ZnO nanowires using micellar gold nanoparticles as catalyst templates". Nanoscale 5, n.º 15 (2013): 7046. http://dx.doi.org/10.1039/c3nr01938a.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
39

Quentel, François, Guillaume Passard y Frederic Gloaguen. "A Binuclear Iron-Thiolate Catalyst for Electrochemical Hydrogen Production in Aqueous Micellar Solution". Chemistry - A European Journal 18, n.º 42 (11 de septiembre de 2012): 13473–79. http://dx.doi.org/10.1002/chem.201201884.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
40

Broxton, Trevor J. y Robin A. Coa. "Micellar catalysis of organic reactions. Part 33. Amide hydrolysis in neutral solution in the presence of a copper-containing micelle". Canadian Journal of Chemistry 71, n.º 5 (1 de mayo de 1993): 670–73. http://dx.doi.org/10.1139/v93-090.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The hydrolysis of 5-nitro-2-(trifluoroacetylamino)benzoic acid (1) has been studied at pH 7 in water and in the presence of micelles of cetyltrimethylammonium bromide (ctab) and of copper-containing micelles formed from the reaction of N,N,N′-trimethyl-N′-hexadecylethylenediamine and cupric chloride. It has been found that the hydrolysis of 1 is inhibited by micelles of ctab but strongly catalysed by the copper-containing micelle at this pH. At a higher pH where the hydroxide ion reaction becomes important the reaction is catalysed by micelles of ctab as well, but the catalysis is stronger by the copper-containing micelle. The effect of added sodium chloride on the rate of reaction is shown to be larger for reaction in the presence of ctab than for reaction in the presence of the copper micelles. Also reported are the effects of the buffer concentration on the rate of reaction at various pH for both micelles. It is concluded that the mechanism of reaction in the copper-containing micelle involves a metal-bound hydroxyl rather than a free hydroxide ion loosely associated with the cationic micelle surface. It is interesting that the catalysis of this reaction by the copper-containing micelle is large enough to allow amide hydrolysis at a reasonable rate at neutral pH at ambient temperature.
41

Nabid, Mohammad Reza y Yasamin Bide. "H40-PCL-PEG unimolecular micelles both as anchoring sites for palladium nanoparticles and micellar catalyst for Heck reaction in water". Applied Catalysis A: General 469 (enero de 2014): 183–90. http://dx.doi.org/10.1016/j.apcata.2013.09.016.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
42

Kumar, Dileep, Malik Abdul Rub y Abdullah M. Asiri. "Synthesis and characterization of geminis and implications of their micellar solution on ninhydrin and metal amino acid complex". Royal Society Open Science 7, n.º 7 (julio de 2020): 200775. http://dx.doi.org/10.1098/rsos.200775.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
In our study, three gemini dicationic surfactants with different methylene group spacer (16-6-16, 16-5-16 and 16-4-16) have been synthesized and characterized in solution by 1 H NMR spectroscopic technique. The implications of gemini micellar solution on ninhydrin and metal amino acid complex ([Cu(II)-Trp] + ) were performed by the means of single-beam UV–visible spectroscopy. The absorbance was noted at regular time intervals and values of rate constant ( k ψ ) were determined by using a computer-based program. Synthesized surfactants proved as an efficient catalyst on the interaction of ninhydrin with metal amino acid complex as compared with conventional surfactant and aqueous systems. The required description regarding the implications of gemini dicationic surfactants are provided in the text in detail. The conductivity technique was applied in order to get critical micelle concentration (cmc) of geminis in the presence and absence of reactants. Catalytic results developed in gemini dicationic surfactant system were explained effectively by pseudo-phase model. Various thermodynamic quantities, viz ., activation energy, E a , activation enthalpy, Δ H # , and activation entropy, Δ S # , were obtained on interaction of ninhydrin with [Cu(II)-Trp] + in gemini systems by applying Eyring equation. A detailed explanation about these evaluated parameters was also made.
43

Poša, Mihalj. "Self-Association of the Anion of 7-Oxodeoxycholic Acid (Bile Salt): How Secondary Micelles Are Formed". International Journal of Molecular Sciences 24, n.º 14 (24 de julio de 2023): 11853. http://dx.doi.org/10.3390/ijms241411853.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
Bile acid anions are steroidal biosurfactants that form primary micelles due to the hydrophobic effect. At higher concentrations of some bile acid anions, secondary micelles are formed; hydrogen bonds connect primary micelles. Monoketo derivatives of cholic acid, which have reduced membrane toxicity, are important for biopharmaceutical examinations. The main goal is to explain why the processes of formation of primary and secondary micelles are separated from each other, i.e., why secondary micelles do not form parallel to primary micelles. The association of the anion of 7-oxodeoxycholic acid (a monoketo derivative of cholic acid) is observed through the dependence of the spin–lattice relaxation time on total surfactant concentration T1 = f(CT). On the function T1 = f(CT), two sharp jumps of the spin–lattice relaxation time are obtained, i.e., two critical micellar concentrations (CMC). The aggregation number of the micelle at 50 mM total concentration of 7-oxodeoxycholic acid anions in the aqueous solution is 4.2 ± 0.3, while at the total concentration of 100 mM the aggregation number is 9.0 ± 0.9. The aggregation number of the micelle changes abruptly in the concentration interval of 80–90 mM (the aggregation number determined using fluorescence measurements). By applying Le Chatelier’s principle, the new mechanism of formation of secondary micelles is given, and the decoupling of the process of formation of primary and secondary micelles at lower concentrations of monomers (around the first critical micellar concentration) and the coupling of the same processes at higher equilibrium concentrations of monomers (around the second critical micellar concentration) is explained. Stereochemically and thermodynamically, a direct mutual association of primary micelles is less likely, but monomeric units are more likely to be attached to primary micelles, i.e., 7-oxodeoxycholic acid anions.
44

Venkateswaran, Krishnan, Mary V. Barnabas, Bill W. Ng y David C. Walker. "Residence-time of muonium at micelles: Effect of added micelles on the reactivity of muonium towards ionic solutes in water". Canadian Journal of Chemistry 66, n.º 8 (1 de agosto de 1988): 1979–83. http://dx.doi.org/10.1139/v88-319.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
Resumen
The effective rate constant for the reaction of muonium with NO3−, S2O32−, and Tl+ ions in water is altered by the addition of micelles. There is a decrease when the charge on the micelle is the same as that of the solute and an increase when their charges are opposite. From the magnitude of the effect a mean residence-time for muonium of 2 ns has been deduced for dodecyl sulphate micelles. This suggests there is barely any preferred localization, because 2 ns is smaller, even, than the expected diffusion time if the micelle core is as viscous as reported. This use of muonium atoms to probe the dynamics of micelles seems to support the view that there are regions of low microviscosity and considerable water penetration within the micellar structure.
45

GOTO, Koichi, Jiro OKAI, Yumiko EJIMA, Takatoshi ITO, Hideo OKAI y Ryuichi UEOKA. "Remarkably Enhanced Enantioselective Hydrolysis of Amino Acid Esters With Tripeptide Catalyst in Micellar Systems." NIPPON KAGAKU KAISHI, n.º 5 (1995): 351–57. http://dx.doi.org/10.1246/nikkashi.1995.351.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
46

Lipshutz, Bruce H. y Subir Ghorai. "PQS: A New Platform for Micellar Catalysis. RCM Reactions in Water, with Catalyst Recycling". Organic Letters 11, n.º 3 (5 de febrero de 2009): 705–8. http://dx.doi.org/10.1021/ol8027829.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
47

Shrikhande, Janhavi J., Manoj B. Gawande y Radha V. Jayaram. "A catalyst-free N-benzyloxycarbonylation of amines in aqueous micellar media at room temperature". Tetrahedron Letters 49, n.º 32 (agosto de 2008): 4799–803. http://dx.doi.org/10.1016/j.tetlet.2008.05.010.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
48

Hájek, Martin, Aleš Vávra, František Skopal, Anna Straková y Miroslav Douda. "The description of catalyst behaviour during transesterification of rapeseed oil – Formation of micellar emulsion". Renewable Energy 159 (octubre de 2020): 938–43. http://dx.doi.org/10.1016/j.renene.2020.06.082.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
49

Cheng, Mingxing, Tian Shi, Hongyu Guan, Shengtian Wang, Xiaohong Wang y Zijiang Jiang. "Clean production of glucose from polysaccharides using a micellar heteropolyacid as a heterogeneous catalyst". Applied Catalysis B: Environmental 107, n.º 1-2 (agosto de 2011): 104–9. http://dx.doi.org/10.1016/j.apcatb.2011.07.002.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.
50

Sobhani, Sara y Zohre Zeraatkar. "A new magnetically recoverable heterogeneous palladium catalyst for phosphonation reactions in aqueous micellar solution". Applied Organometallic Chemistry 30, n.º 1 (26 de octubre de 2015): 12–19. http://dx.doi.org/10.1002/aoc.3392.

Texto completo
Los estilos APA, Harvard, Vancouver, ISO, etc.

Pasar a la bibliografía