Littérature scientifique sur le sujet « Surfactant Interaction »

Créez une référence correcte selon les styles APA, MLA, Chicago, Harvard et plusieurs autres

Choisissez une source :

Consultez les listes thématiques d’articles de revues, de livres, de thèses, de rapports de conférences et d’autres sources académiques sur le sujet « Surfactant Interaction ».

À côté de chaque source dans la liste de références il y a un bouton « Ajouter à la bibliographie ». Cliquez sur ce bouton, et nous générerons automatiquement la référence bibliographique pour la source choisie selon votre style de citation préféré : APA, MLA, Harvard, Vancouver, Chicago, etc.

Vous pouvez aussi télécharger le texte intégral de la publication scolaire au format pdf et consulter son résumé en ligne lorsque ces informations sont inclues dans les métadonnées.

Articles de revues sur le sujet "Surfactant Interaction"

1

Cheng, Chao, et Shi-Yong Ran. « Interaction between DNA and Trimethyl-Ammonium Bromides with Different Alkyl Chain Lengths ». Scientific World Journal 2014 (2014) : 1–9. http://dx.doi.org/10.1155/2014/863049.

Texte intégral
Résumé :
The interaction betweenλ—DNA and cationic surfactants with varying alkyl chain lengths was investigated. By dynamic light scattering method, the trimethyl-ammonium bromides-DNA complex formation was shown to be dependent on the length of the surfactant’s alkyl chain. For surfactants with sufficient long alkyl chain (CTAB, TTAB, DTAB), the compacted particles exist with a size of ~60–110 nm at low surfactant concentrations. In contrast, high concentration of surfactants leads to aggregates with increased sizes. Atomic force microscope scanning also supports the above observation. Zeta potential measurements show that the potential of the particles decreases with the increase of surfactant concentration (CTAB, TTAB, DTAB), which contributes much to the coagulation of the particles. For OTAB, the surfactant with the shortest chain in this study, it cannot fully neutralize the charges of DNA molecules; consequently, the complex is looser than other surfactant-DNA structures.
Styles APA, Harvard, Vancouver, ISO, etc.
2

Taba, Paulina, Russell F. Howe et Graine Moran. « FTIR AND NMR STUDIES OF ADSORBED CETHYLTRIMETHYLAMMONIUM CHLORIDE IN MCM-41 MATERIALS ». Indonesian Journal of Chemistry 8, no 1 (17 juin 2010) : 1–6. http://dx.doi.org/10.22146/ijc.21639.

Texte intégral
Résumé :
The high use of surface-active agents (surfactants) by industry and households today leads to environmental pollution, therefore treatments are required to remove such substances from the environment. One of the important and widely used methods for removal of substances from solution is adsorption. In this research, MCM-41 and its modified product of MCM41-TMCS were used to adsorb cationic surfactants, cethyltrimethylammonium chloride, CTAC. FTIR and NMR methods were used to study the interaction between the surfactants and the adsorbents. MCM-41 was synthesized hydrothermally at 100 oC and its modification was conducted by silylation of MCM-41 with trimethylchloro silane (MCM41-TMCS). Both unmodified and modified MCM-41 can adsorb the surfactant. The interaction of CTAC with MCM-41 was mostly the electrostatic interaction between the electropositive end of the surfactant and MCM-41, whereas in modified MCM-41 hydrophobic interactions become more dominant. These hydrophobic interactions appear however to involve the methyl groups on the head group of the surfactant interacting with the modified surface. Keywords: FTIR, NMR, adsorbed CTAC, MCM-41 materials
Styles APA, Harvard, Vancouver, ISO, etc.
3

Yang, Jia, et Rajinder Pal. « Investigation of Surfactant-Polymer Interactions Using Rheology and Surface Tension Measurements ». Polymers 12, no 10 (8 octobre 2020) : 2302. http://dx.doi.org/10.3390/polym12102302.

Texte intégral
Résumé :
The interactions between surfactants and a drag-reducing polymer were investigated at a low polymer concentration of 500 ppm, using measurements of the rheology and surface activity of surfactant-polymer solutions. A well-known drag-reducing polymer (anionic sodium carboxymethyl cellulose) and five different surfactants (two anionic, two non-ionic, and one zwitterionic) were selected for the interaction studies. The surfactant-polymer solutions were shear thinning in nature, and they followed the power law model. The interaction between the surfactant and polymer had a strong effect on the consistency index of the solution and a marginal effect on the flow behavior index. The surface tension versus surfactant concentration plots were interpreted in terms of the interactions between surfactant and polymer. The critical aggregation concentration (CAC) of the surfactant was estimated based on the surface tension and rheological data. The CAC values of the same charge surfactants as that of the polymer were found to be significantly higher than other combinations of surfactant and polymer, such as non-ionic surfactant/anionic polymer, and zwitterionic surfactant/anionic polymer.
Styles APA, Harvard, Vancouver, ISO, etc.
4

Lai, Chiu-Chun, Kuo-Shien Huang, Po-Wei Su, Chang-Mou Wu et Ching-Nan Huang. « Interactions of modified Gemini surfactants : Interactions with direct dyes and dyeing properties in cotton fabrics ». Modern Physics Letters B 33, no 14n15 (28 mai 2019) : 1940002. http://dx.doi.org/10.1142/s0217984919400025.

Texte intégral
Résumé :
This study investigated dye–surfactant interactions between a series of modified Gemini surfactants and commercial direct dyes in aqueous solution and their corresponding effects on cotton fabric dyeing. A surface tension meter was also used to measure surface activities of compounds containing electrolyte under conditions similar to those in dyeing processes. The surface tension measurements showed lower than normal surface tension in surfactant solutions containing electrolyte. From the UV-Vis spectra, the isosbestic point indicated that dye–surfactant complexes had formed and existed as hydrophilic interaction between direct dyes and modified Gemini surfactants. When dyeing cotton fabric with red dye and orange dye, the presence of these surfactants decreased dye uptake rate but increased for blue dye because the dye–surfactant interaction had formed a hydrophilic complex.
Styles APA, Harvard, Vancouver, ISO, etc.
5

Ostos, Francisco José, José Antonio Lebrón, María Luisa Moyá, Eva Bernal, Ana Flores, Cristian Lépori, Ángeles Maestre, Francisco Sánchez, Pilar López-Cornejo et Manuel López-López. « Potentiometric Study of Carbon Nanotube/Surfactant Interactions by Ion-Selective Electrodes. Driving Forces in the Adsorption and Dispersion Processes ». International Journal of Molecular Sciences 22, no 2 (15 janvier 2021) : 826. http://dx.doi.org/10.3390/ijms22020826.

Texte intégral
Résumé :
The interaction (adsorption process) of commercial ionic surfactants with non-functionalized and functionalized carbon nanotubes (CNTs) has been studied by potentiometric measurements based on the use of ion-selective electrodes. The goal of this work was to investigate the role of the CNTs’ charge and structure in the CNT/surfactant interactions. Non-functionalized single- (SWCNT) and multi-walled carbon nanotubes (MWCNT), and amine functionalized SWCNT were used. The influence of the surfactant architecture on the CNT/surfactant interactions was also studied. Surfactants with different charge and hydrophobic tail length (sodium dodecyl sulfate (SDS), octyltrimethyl ammonium bromide (OTAB), dodecyltrimethyl ammonium bromide (DoTAB) and hexadecyltrimethyl ammonium bromide (CTAB)) were studied. According to the results, the adsorption process shows a cooperative character, with the hydrophobic interaction contribution playing a key role. This is made evident by the correlation between the free surfactant concentration (at a fixed [CNT]) and the critical micellar concentration, cmc, found for all the CNTs and surfactants investigated. The electrostatic interactions mainly determine the CNT dispersion, although hydrophobic interactions also contribute to this process.
Styles APA, Harvard, Vancouver, ISO, etc.
6

LIU, HAO-YANG, XIAN-WU ZOU, YIN-QUAN YUAN et ZHUN-ZHI JIN. « EFFECTS OF INTERACTION WITH SOLVENT AND CHAIN CONFORMATION OF SURFACTANTS ON EMULSIFICATION ». Modern Physics Letters B 15, no 24 (20 octobre 2001) : 1061–68. http://dx.doi.org/10.1142/s0217984901002853.

Texte intégral
Résumé :
The details of the emulsification process has been investigated by discontinuous molecular dynamic simulation. The surfactants help to bring about emulsification. The emulsification can be divided crudely into two stages: splitting and uniting process. The splitting and uniting of oil droplets occurs in this position, where surfactants at the interface is rather scarce. The effects of the conformation of surfactant chain and the strength of surfactant–water and surfactant–oil interactions on emulsification were also studied. The surfactants with longer tail and stronger surfactant–water and surfactant–oil interactions promote the emulsification more.
Styles APA, Harvard, Vancouver, ISO, etc.
7

Hosseinzadeh, Reza, Mohammad Gheshlagi, Rahele Tahmasebi et Farnaz Hojjati. « Spectrophotometric study of interaction and solubilization of procaine hydrochloride in micellar systems ». Open Chemistry 7, no 1 (1 mars 2009) : 90–95. http://dx.doi.org/10.2478/s11532-008-0078-4.

Texte intégral
Résumé :
AbstractThe interaction of Procaine hydrochloride (PC) with cationic, anionic and non-ionic surfactants; cetyltrimethylammonium bromide (CTAB), sodium dodecyl sulfate (SDS) and triton X-100, were investigated. The effect of ionic and non-ionic micelles on solubilization of Procaine in aqueous micellar solution of SDS, CTAB and triton X-100 were studied at pH 6.8 and 29°C using absorption spectrophotometry. By using pseudo-phase model, the partition coefficient between the bulk water and micelles, Kx, was calculated. The results showed that the micelles of CTAB enhanced the solubility of Procaine higher than SDS micelles (Kx = 96 and 166 for SDS and CTAB micelles, respectively) but triton X-100 did not enhanced the solubility of drug because of weak interaction with Procaine. From the resulting binding constant for Procaine-ionic surfactants interactions (Kb = 175 and 128 for SDS and CTAB surfactants, respectively), it was concluded that both electrostatic and hydrophobic interactions affect the interaction of surfactants with cationic procaine. Electrostatic interactions have a great role in the binding and consequently distribution of Procaine in micelle/water phases. These interactions for anionic surfactant (SDS) are higher than for cationic surfactant (CTAB). Gibbs free energy of binding and distribution of procaine between the bulk water and studied surfactant micelles were calculated.
Styles APA, Harvard, Vancouver, ISO, etc.
8

Tazhibayeva, Sagdat, Kuanyshbek Musabekov, Zhenis Kusainova, Ardak Sapieva et Nurlan Musabekov. « Complex Formation of Polyacrylic Acid with Surfactants of Different Hydrophobicity ». Applied Mechanics and Materials 752-753 (avril 2015) : 212–16. http://dx.doi.org/10.4028/www.scientific.net/amm.752-753.212.

Texte intégral
Résumé :
Complex formation processes of polyelectrolytes with surfactant ions are close model to protein - lipid interactions in living organisms. Furthermore, polymer – surfactant complexes are widely used as stabilizers of industrial dispersions and structurants of soil. When using the polymer-surfactant complexes the hydrophilic-lipophilic balance has the great importance. The interaction of polyacrylic acid with alkylammonium salts of different hydrophobicity: cetyltrimethylammonium bromide, dilaurildimethylammonium bromide and dioctadecyldimethylammonium chloride was studied by potentiometry, spectrophotometry, viscometry and electrophoresis methods. It was established that the complex formation of polyacrylic acid with cationic surfactants is carried out due to the electrostatic interaction between carboxyl groups of the polymer and cations of surfactants, which stabilized by hydrophobic interactions between their non-polar parts. The phenomenon of hysteresis in the change of the reduced viscosity of system surfactant /polyacrylic acid with temperature variation in the range of 20-60 °C was found. The possibility of using the complex formation process for water purification from CTAB has been shown. The degree of purification is 99.6-99.8%.
Styles APA, Harvard, Vancouver, ISO, etc.
9

Reddy, M. C. Somasekhara, S. M. Sarvar Jahan, K. Sridevi et G. V. Subba Reddy. « Investigations on Natural Surfactant obtained from Soap-Nuts through Spectrophotometric Interactions with Congo Red and Comparison with Commercial Surfactants ». Asian Journal of Chemistry 31, no 4 (27 février 2019) : 907–16. http://dx.doi.org/10.14233/ajchem.2019.21849.

Texte intégral
Résumé :
A natural surfactant (NS) of plant-base was obtained from the fruit pericarp of soapnuts by using a simple and economical method. The interaction of this natural surfactant with direct dye, anionic dye, Congo red (CR) was studied spectrophotometrically in sub-micellar and micelle concentration range in aqueous solution. These interactions (CR-NS) were compared with that of CR-CTAB (cationic surfactant-cetyl trimethylammonium bromide, CTAB), CR-SDS (anionic surfactant-sodium dodecyl sulphate, SDS) and CR-TX 100 (neutral surfactant - Triton X-100, TX 100) and were useful to understand the nature of natural surfactant. The mechanism of formation of complex due to interactions between Congo red and natural surfactant was suggested. This spectrophotometric method was used for the determination of critical micelle concentration (CMC), at which the formation of micelles was started. The CMC values obtained spectrophotometrically for the natural surfactant was coincided with the experimental value available in the literature. A definite change in the absorbance maxima of Congo red in the presence of natural surfactant (micelles of natural surfactant) was also observed. The change in maxima was also interpreted in terms of pH and CMC. The equilibrium constant of interaction between Congo red and natural surfactant was calculated on the theoretical model. The stability of the complexes of Congo red with different surfactants like CTAB, SDS, TX 100 and natural surfactant may be written in increasing order as: CR-TX 100 > CR-CTAB > CR-NS > CR-SDS. The biodegradable, non-toxic, inexpensive, environmental friendly, renewable natural surfactant was suggested in place of synthetic surfactants.
Styles APA, Harvard, Vancouver, ISO, etc.
10

Nazarova, Anastasia, Arthur Khannanov, Artur Boldyrev, Luidmila Yakimova et Ivan Stoikov. « Self-Assembling Systems Based on Pillar[5]arenes and Surfactants for Encapsulation of Diagnostic Dye DAPI ». International Journal of Molecular Sciences 22, no 11 (3 juin 2021) : 6038. http://dx.doi.org/10.3390/ijms22116038.

Texte intégral
Résumé :
In this paper, we report the development of the novel self-assembling systems based on oppositely charged Pillar[5]arenes and surfactants for encapsulation of diagnostic dye DAPI. For this purpose, the aggregation behavior of synthesized macrocycles and surfactants in the presence of Pillar[5]arenes functionalized by carboxy and ammonium terminal groups was studied. It has been demonstrated that by varying the molar ratio in Pillar[5]arene-surfactant systems, it is possible to obtain various types of supramolecular systems: host–guest complexes at equimolar ratio of Pillar[5]arene-surfactant and interpolyelectrolyte complexes (IPECs) are self-assembled materials formed in aqueous medium by two oppositely charged polyelectrolytes (macrocycle and surfactant micelles). It has been suggested that interaction of Pillar[5]arenes with surfactants is predominantly driven by cooperative electrostatic interactions. Synthesized stoichiometric and non-stoichiometric IPECs specifically interact with DAPI. UV-vis, luminescent spectroscopy and molecular docking data show the structural feature of dye-loaded IPEC and key role of the electrostatic, π–π-stacking, cation–π interactions in their formation. Such a strategy for the design of supramolecular Pillar[5]arene-surfactant systems will lead to a synergistic interaction of the two components and will allow specific interaction with the third component (drug or fluorescent tag), which will certainly be in demand in pharmaceuticals and biomedical diagnostics.
Styles APA, Harvard, Vancouver, ISO, etc.

Thèses sur le sujet "Surfactant Interaction"

1

Gomez, Gil Leticia. « The interaction between cholesterol and surfactant protein-c in lung surfactant ». Doctoral thesis, Universite Libre de Bruxelles, 2009. http://hdl.handle.net/2013/ULB-DIPOT:oai:dipot.ulb.ac.be:2013/210205.

Texte intégral
Résumé :
The presence of cholesterol is critical in defining a dynamic lateral structure in pulmonary

surfactant membranes, including the segregation of fluid-ordered and fluid-disordered phases.

However, an excess of cholesterol has been associated with impaired surface activity both in

surfactant models and in surfactant from injured lungs. It has also been reported that surfactant

protein SP-C interacts with cholesterol in lipid/protein interfacial films. In the present study, we

have analyzed the effect of SP-C on the thermodynamic properties of phospholipid membranes

containing cholesterol and on the ability of lipid/protein complexes containing surfactant

proteins and cholesterol to form and re-spread interfacial films capable of producing very low

surface tensions upon repetitive compression-expansion cycling. We have also analyzed the effect of cholesterol on the

structure, orientation and dynamic properties of SP-C embedded in physiologically relevant

model membranes.


Doctorat en Sciences
info:eu-repo/semantics/nonPublished

Styles APA, Harvard, Vancouver, ISO, etc.
2

Farnoud, Amir Mohammad. « Interaction of polymeric particles with surfactant interfaces ». Diss., University of Iowa, 2013. https://ir.uiowa.edu/etd/4627.

Texte intégral
Résumé :
Films of phospholipids and biologically relevant surfactants at the air-water interface provide a well-defined medium to study molecular alignment, phase behavior and interactions of biomembranes and lung surfactant with exogenous materials. Interactions between lung surfactant interfaces and solid particles are of particular interest due to the increased use of nanomaterials in industrial applications and the promise of polymeric particles in pulmonary drug delivery. Understanding such interactions is necessary to avoid potential adverse effects on surfactant function after exposure to particles. In this thesis, the mechanisms of surfactant inhibition after exposure to submicron particles via different routes were investigated. The effects of carboxyl-modified polystyrene particles (200 nm) on films of dipalmitoyl phosphatidylcholine (DPPC) and Infasurf (calf lung surfactant extract) were studied. Surfactants were exposed to different concentrations of particles in a Langmuir trough with symmetric surface compression and expansion. Surface tension, potential, microstructure and topology were examined to monitor particle effects on surfactant function. Several methods of surfactant exposure to particles were studied: particle injection into the subphase after spreading surfactant monolayers (subphase injection), mixing the particles with the subphase and spreading the surfactant on top (monolayer addition) and particle aerosolization onto surfactant films. Studies with DPPC monolayers revealed that particle-surfactant interactions are dependent on the particle introduction method. In the subphase injection method, particles did not penetrate the monolayer and no inhibitory effects on surfactant function were observed. However, in the monolayer addition method, particles caused a premature monolayer collapse and hindered surfactant respreading likely by penetrating into the DPPC monolayer. Finally, particle aerosolization on surfactant was performed to mimic the physiologically relevant route of surfactant exposure to particles. Particle aerosolization on DPPC monolayers significantly inhibited surfactant function in the lung-relevant surface tension range. When aerosolized on Infasurf, particles caused inhibitory effects as a function of time suggesting adsorption of surfactant components on particle surfaces as the main mechanism of interaction. This research will enhance understanding of the mechanisms of particle-induced surfactant dysfunction, thereby providing information for the safe design of polymeric particles for drug delivery and for developing guidelines for particles used in occupational settings.
Styles APA, Harvard, Vancouver, ISO, etc.
3

Windsor, Rosemary. « Polymer surfactant interaction studied by sum frequency spectroscopy ». Thesis, University of Cambridge, 2002. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.620464.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
4

Carnell, Sarah. « Surfactant interaction and persistence at the contact lens surface ». Thesis, Aston University, 2015. http://publications.aston.ac.uk/37488/.

Texte intégral
Résumé :
The interaction of poloxamer and poloxamine (Pluronic® and Tetronic®) surfactants with hydrogel and silicone hydrogel contact lens surfaces is of interest for this thesis. The persistence of surfactant molecules at the lens surface can indicate how long the surface has been modified. It is therefore important to observe and characterise the surface and surfactant behaviour separately. Characterisation of the contact lenses was carried out through dehydrated sessile drop measurements and surface energy calculations. Silicone-containing materials tended to be most hydrophobic regardless of water content. Static and dynamic surface tension measurements were used to assess the surfactants and the critical micelle concentration was also observed. Pluronics® and Tetronics® do not behave as simple low molecular weight surfactants; their structure and size mean they are less mobile in solution and may be able to form mono molecular micelles. Surfactants with different molecular structure, molecular weight and hydrophobicity were used to observe how these properties affect surface tension behaviour and influence surfactant persistence. The aim of the work was observe the persistence of surfactants at the lens surface, any difference between the surfactant persistence, and the possibility to predict surfactant persistence on a lens. The ex vivo work presented here shows little distinction between surface tension measurements over time or between treated and untreated materials. It is not possible to measure in vivo surfactant persistence with surface tension techniques and therefore necessary to create in vitro models to assess surfactant behaviour. A simplified in vitro eye model was created to assess preliminary observations. These results and observations were used to progressively alter the model and create a more ‘eye-like’ system. Large hydrophobic Tetronics® were most persistent at the lens surface; hydrophobic drive was considered the most influential factor. In addition to this, the contact lens material and condition prior to surfactant treatment also had an effect on persistence. Materials containing PVP showed increased surfactant persistence, which was increased further when the lenses were dehydrated prior to surfactant treatment. Lens dehydration had no effect on persistence if PVP was not present in the lens material.
Styles APA, Harvard, Vancouver, ISO, etc.
5

McKenzie, Zofi. « Nanotoxicology : nanoparticle interaction with surfactant proteins A and D ». Thesis, University of Southampton, 2013. https://eprints.soton.ac.uk/390356/.

Texte intégral
Résumé :
Numerous epidemiological and toxicological studies have associated enhanced exposure to ambient air pollution with reduced resolution and increased incidence of respiratory infections. Surfactant Proteins A (SP-A) and SP-D are innate immune molecules within the lung and are important mediators in the resolution and clearance of microbial infections. They have also been implicated in the opsonisation and clearance of inorganic particulates in vitro. This study aimed to investigate the interaction of SP-A and SP-D with model 100nm unmodified (U-PS) and amine modified polystyrene (A-PS) nanoparticles. Firstly, it was hypothesised that the particle interaction with these proteins would alter particle clearance by macrophages and secondly that the sequestration of SP-A and SP-D by particles would result in a reduction in the anti-microbial function of these proteins. SP-A and SP-D were purified from the bronchoalveolar lavage fluid of subjects with alveolar proteinosis. Using absorption, turbidity, size and zeta potential measurements SP-A and SP-D were shown to interact with A-PS and U-PS particles and the extent of these interactions were dependent on the zeta potential of the particles. SP-A and SP-D altered the colloidal stability of the particles and this was related to the effect of each protein on the differential particle uptake by macrophages. In vitro influenza A virus (IAV) infection models were optimised using flow cytometry to detect surfactant protein mediated neutralisation of this virus at sub-maximal levels in cell lines representing cells found within the alveolus. These models were used to study the effect of U-PS and A-PS particles on surfactant protein mediated neutralisation of IAV. The results showed that nanoparticles can modulate the vitro function of SP-A and SP-D in a biphasic fashion in alveolar epithelial cells. However, this effect was dependent on a number of factors, including the particle, the protein and cell type under investigation. The identification of unlabelled lipids and nanoparticles in vitro by coherent anti-stokes raman scattering (CARS) was also be discussed.
Styles APA, Harvard, Vancouver, ISO, etc.
6

Hohenschutz, Max. « Nano-ions in interaction with non-ionic surfactant self-assemblies ». Thesis, Montpellier, 2020. http://www.theses.fr/2020MONTS064.

Texte intégral
Résumé :
Les ions de taille nanométrique (nano-ions), tels que les clusters ioniques de bore, les polyoxométalates (POM) et les grands ions organiques, ont suscité un intérêt remarquable ces dernières années en raison de leur capacité à s’adsorber ou se lier à des systèmes chimiques électriquement neutres, tels que les molécules hôtes macrocycliques, les nanoparticules, les tensioactifs et les polymères, etc. Il a été démontré que ces processus d'adsorption ou de liaison sont induits par un phénomène médié par solvant, l'effet chaotropique, qui pousse le nano-ion de la masse d'eau vers une interface. Ainsi, l'eau d'hydratation de l'ion et de l'interface est libérée dans la masse d'eau, ce qui entraîne une restitution de la structure intrinsèque de l'eau. Cet effet est particulièrement fort pour les nano-ions. Ils sont par conséquent appelés ions superchaotropiques ou hydrophobes dans le prolongement des ions classiques (faiblement) chaotropiques tels que le SCN-. Tous les superchaotropes couramment étudiés, bien que chimiquement divers, partagent des caractéristiques physiques telles qu'une faible densité de charge et une grande polarisabilité. Les effets des nano-ions sur les auto-assemblages de tensioactifs non ioniques éthoxylés, les phases micellaires et bicouches, sont ici élucidés pour tirer des conclusions sur leur nature chaotropique et/ou hydrophobe. En combinant la diffusion aux petits angles des neutrons et des rayons X (SANS et SAXS), et les diagrammes de phase, les systèmes tensioactifs non ioniques/nano-ion sont examinés et comparés, du nanomètre à l'échelle macroscopique. Ainsi, il est montré que tous les nano-ions étudiés induisent un chargement électrique des assemblages de tensioactifs ainsi qu'une déshydratation des têtes de tensioactif non-ionique. En outre, les ions chaotropiques ou hydrophobes diffèrent dans leurs effets sur la forme micellaire. Les ions chaotropiques entraînent les micelles allongées de tensioactif non-ionique vers les micelles sphériques (augmentation de la courbure), tandis que les ions hydrophobes provoquent une transition vers les phases bicouches (diminution de la courbure). Il est conclu que les nano-ions superchaotropiques agissent comme des tensioactifs ioniques car leur ajout à des systèmes de tensioactifs non ioniques provoque un effet de charge. Cependant, les nano-ions et les tensioactifs ioniques sont fondamentalement différents par leur association avec l'ensemble des tensioactifs non ioniques. Le nano-ion s'adsorbe sur les têtes des tensioactifs non ioniques par effet chaotropique, tandis que le tensioactif ionique s'ancre dans les micelles entre les queues des tensioactifs non ioniques par effet hydrophobe. La comparaison des effets de l'ajout de nano-ions ou de tensioactifs ioniques à des tensioactifs non ioniques a été approfondie sur les mousses. Les mousses ont été étudiées en ce qui concerne l'épaisseur du film de mousse, le drainage dans le temps et la stabilité, respectivement en utilisant la SANS, l'analyse d'image et la conductométrie. Le POM superchaotropique testé (SiW12O404-, SiW) ne mousse pas dans l'eau contrairement au SDS classique de tensioactif ionique. Néanmoins, l'ajout de petites quantités de SiW ou de SDS à une solution moussante de tensioactif non ionique a permis d'obtenir des mousses plus humides avec une durée de vie plus longue. Entre-temps, l'épaisseur du film de mousse (déterminée en SANS) est augmentée en raison de la charge électrique des monocouches de tensioactifs non ioniques dans le film de mousse. Il est conclu que le comportement remarquable des nano-ions - ici sur les systèmes tensioactifs non ioniques - peut être étendu aux systèmes colloïdaux, tels que les mousses, les polymères, les protéines ou les nanoparticules. Cette thèse démontre que le comportement superchaotropique des nano-ions est un outil polyvalent qui peut être utilisé dans de nouvelles formulations de matériaux et d'applications de la matière molle
Nanometer-sized ions (nano-ions), such as ionic boron clusters, polyoxometalates (POMs) and large organic ions, have spawned remarkable interest in recent years due to their ability to adsorb or bind to electrically neutral chemical systems, such as macrocyclic host molecules, colloidal nano-particles, surfactants and polymers etc. The underlying adsorption or binding processes were shown to be driven by a solvent-mediated phenomenon, the chaotropic effect, which drives the nano-ion from the water bulk towards an interface. Thus, hydration water of the ion and the interface is released into the bulk resulting in a bulk water structure recovery. This effect is particularly strong for nano-ions. Therefore, they were termed superchaotropic or hydrophobic ions as an extension to classical (weakly) chaotropic ions such as SCN-. All commonly studied superchaotropes, though chemically diverse, share physical characteristics such as low charge density and high polarizability. Herein, the effects of nano-ions on ethoxylated non-ionic surfactant self-assemblies, micellar and bilayer phases, are elucidated to draw conclusions on their chaotropic and/or hydrophobic nature. By combining small angle scattering of neutrons and x-rays (SANS and SAXS), and phase diagrams, non-ionic surfactant/nano-ion systems are examined and compared, from the nanometer to the macroscopic scale. Thus, all studied nano-ions are found to induce a charging of the surfactant assemblies along with a dehydration of the non-ionic surfactant head groups. Furthermore, chaotropic and hydrophobic ions differ in their effects on the micellar shape. Superchaotropic ions drive the elongated non-ionic surfactant micelles towards spherical micelles (increase in curvature), whereas hydrophobic ions cause a transition towards bilayer phases (decrease in curvature). It is concluded that superchaotropic nano-ions act like ionic surfactants because their addition to non-ionic surfactant systems causes a charging effect. However, nano-ions and ionic surfactants are fundamentally different by their association with the non-ionic surfactant assembly. The nano-ion adsorbs to the non-ionic surfactant heads by the chaotropic effect, while the ionic surfactant anchors into the micelles between the non-ionic surfactant tails by the hydrophobic effect. The comparison of the effects of adding nano-ions or ionic surfactant to non-ionic surfactant was further investigated on foams. The foams were investigated regarding foam film thickness, drainage over time and stability, respectively using SANS, image analysis and conductometry. The tested superchaotropic POM (SiW12O404-, SiW) does not foam in water in contrast to the classical ionic surfactant SDS. Nevertheless, addition of small amounts of SiW or SDS to a non-ionic surfactant foaming solution resulted in wetter foams with longer lifetimes. Meanwhile, the foam film thickness (determined in SANS) is increased due to the electric charging of the non-ionic surfactant monolayers in the foam film. It is concluded that the remarkable behavior of nano-ions – herein on non-ionic surfactant systems – can be extended to colloidal systems, such as foams, polymers, proteins or nanoparticles. This thesis demonstrates that the superchaotropic behavior of nano-ions is a versatile tool to be used in novel formulations of soft matter materials and applications
Styles APA, Harvard, Vancouver, ISO, etc.
7

Crichton, Donna. « The interaction of oils with surfactant monolayers at the air-water surface ». Thesis, University of Hull, 1998. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.310247.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
8

Ocampo, Minette C. « Protein-Lipid Interactions with Pulmonary Surfactant Using Atomic Force Microscopy ». The Ohio State University, 2014. http://rave.ohiolink.edu/etdc/view?acc_num=osu1395050693.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
9

Höhn, Sebastian [Verfasser]. « Interaction of Pluronic polymers with sugar surfactant in microemulsions designed for decontamination / Sebastian Höhn ». Bielefeld : Universitätsbibliothek Bielefeld, 2016. http://d-nb.info/1101694106/34.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
10

Nilsson, Peter. « Interaction between Crosslinked Polyelectrolyte Gels and Oppositely Charged Surfactants ». Doctoral thesis, Uppsala University, Department of Pharmacy, 2007. http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-8216.

Texte intégral
Résumé :

The interactions between anionic, crosslinked gels and cationic surfactants have been investigated. When exposed to oppositely charged surfactant, the gel collapses into a dense complex of polyion and micelles. During deswelling, the gel phase separates into a micelle-rich, collapsed surface phase, and a swollen, micelle-free core, both still part of the same network. As more surfactant is absorbed, the surface phase grows at the expense of the core, until the entire gel has collapsed. Polyacrylate (PA) gels with dodecyl- (C12TAB), and cetyltrimethylammonium bromide (C16TAB), as well as hyaluronate gels with cetylpyridinium chloride, have been studied.

Kinetic experiments have been performed on macro- as well as microgels, using micromanipulator assisted light microscopy for the latter. A surfactant diffusion controlled deswelling model has been employed to describe the deswelling. The deswelling kinetics of PA microgels have been shown to be controlled by surfactant diffusion through the stagnant layer surrounding the gel, as the surface phase is relatively thin for the major part of the deswelling. For macroscopic PA gels the surface phase is thicker, and the kinetics with C12TAB were therefore also influenced by diffusion through the surface phase, while for C16TAB they were dominated by it.

Relevant parameters have also been determined using equilibrium experiments. An irregular, balloon-forming deswelling pattern, mainly found for macrogels, as well as unexpectedly long lag times and slow deswelling for microgels, are reported and discussed.

The microstructure of fully collapsed PA/C12TAB complexes has been studied using small-angle X-ray scattering. A cubic Pm3n structure was found at low salt concentration, which melted into a disordered micellar phase as the salt concentration was increased. Further increasing the salt concentration dissolved the micelles, resulting in no ordering.

Styles APA, Harvard, Vancouver, ISO, etc.

Livres sur le sujet "Surfactant Interaction"

1

Martinez-Santiago, Jose. Polyelectrolyte-Surfactant Phase Behavior and Mechanisms of Interaction in Multi-Component Systems. [New York, N.Y.?] : [publisher not identified], 2015.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
2

Dias, Rita, et Bjrn Lindman, dir. DNA Interactions with Polymers and Surfactants. Hoboken, NJ, USA : John Wiley & Sons, Inc., 2008. http://dx.doi.org/10.1002/9780470286364.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
3

1942-, Lindman Björn, et Dias Rita, dir. DNA interactions with polymers and surfactants. Hoboken, N.J : John Wiley, 2008.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
4

1926-, Goddard E. D., et Ananthapadmanabhan Kavssery P. 1952-, dir. Interactions of surfactants with polymers and proteins. Boca Raton : CRC Press, 1993.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
5

Peter, Gehr, dir. Particle-lung interactions. 2e éd. New York : Informa Healthcare, 2010.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
6

Dias, Rita, et Bjorn Lindman. DNA Interactions with Polymers and Surfactants. Wiley & Sons, Incorporated, John, 2008.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
7

Dias, Rita, et Bjorn Lindman. DNA Interactions with Polymers and Surfactants. Wiley-Blackwell, 2008.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
8

Goddard, E. Desmond. Interactions of Surfactants with Polymers and Proteins. Sous la direction de E. D. Goddard et K. P. Ananthapadmanabhan. CRC Press, 2018. http://dx.doi.org/10.1201/9781351073783.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
9

Goddard, E. Desmond. Interactions of Surfactants with Polymers and Proteins. Taylor & Francis Group, 2018.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
10

Goddard, E. Desmond. Interactions of Surfactants with Polymers and Proteins. Taylor & Francis Group, 2018.

Trouver le texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.

Chapitres de livres sur le sujet "Surfactant Interaction"

1

Tadros, Tharwat. « Polyelectrolyte-surfactant Interaction ». Dans Encyclopedia of Colloid and Interface Science, 945. Berlin, Heidelberg : Springer Berlin Heidelberg, 2013. http://dx.doi.org/10.1007/978-3-642-20665-8_129.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
2

Goddard, E. D. « On Polymer/Surfactant Interaction ». Dans Surfactants in Solution, 219–42. Boston, MA : Springer US, 1991. http://dx.doi.org/10.1007/978-1-4615-3836-3_16.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
3

Marangoni, D. Gerrard, Andrew P. Rodenhiser, Jill M. Thomas et Jan C. T. Kwak. « Interaction of Alcohols and Ethoxylated Alcohols with Anionic and Cationic Micelles ». Dans Mixed Surfactant Systems, 194–209. Washington, DC : American Chemical Society, 1992. http://dx.doi.org/10.1021/bk-1992-0501.ch011.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
4

Tanaka, Fumihiko. « Polymer-surfactant interaction in thermoreversible gels ». Dans Molecular Interactions and Time-Space Organization in Macromolecular Systems, 81–89. Berlin, Heidelberg : Springer Berlin Heidelberg, 1999. http://dx.doi.org/10.1007/978-3-642-60226-9_9.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
5

Pan, Xin, Zhengwei Huang et Chuanbin Wu. « Interaction between Inhalable Nanomedicines and Pulmonary Surfactant ». Dans Organ Specific Drug Delivery and Targeting to the Lungs, 109–48. Boca Raton : CRC Press, 2022. http://dx.doi.org/10.1201/9781003182566-5.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
6

Gilányi, T., I. Varga et R. Mészáros. « Molecular interaction model of polymer–surfactant complex formation ». Dans From Colloids to Nanotechnology, 179–83. Berlin, Heidelberg : Springer Berlin Heidelberg, 2004. http://dx.doi.org/10.1007/978-3-540-45119-8_30.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
7

Zhou, Ting, et Guiying Xu. « Aggregation Behavior of Ionic Liquid-Based Gemini Surfactants and Their Interaction with Biomacromolecules ». Dans Ionic Liquid-Based Surfactant Science, 127–49. Hoboken, NJ : John Wiley & Sons, Inc, 2015. http://dx.doi.org/10.1002/9781118854501.ch6.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
8

Pérez-Gil, J., A. Cruz, M. L. F. Ruano, E. Miguel, I. Plasencia et C. Casals. « Interaction of Pulmonary Surfactant-Associated Proteins with Phospholipid Vesicles ». Dans Molecular Dynamics of Biomembranes, 399–420. Berlin, Heidelberg : Springer Berlin Heidelberg, 1996. http://dx.doi.org/10.1007/978-3-642-61126-1_31.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
9

Högberg, Ida, Fredrik Andersson, Erik Hedenström, Magnus Norgren et Håkan Edlund. « The Interaction Parameter in Binary Surfactant Mixtures of a Chelating Surfactant and a Foaming Agent ». Dans Trends in Colloid and Interface Science XXIV, 17–20. Berlin, Heidelberg : Springer Berlin Heidelberg, 2011. http://dx.doi.org/10.1007/978-3-642-19038-4_3.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
10

Mitsuyasu, H., et T. Honda. « The Effects of Surfactant on Certain Air—Sea Interaction Phenomena ». Dans Wave Dynamics and Radio Probing of the Ocean Surface, 95–115. Boston, MA : Springer US, 1986. http://dx.doi.org/10.1007/978-1-4684-8980-4_6.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.

Actes de conférences sur le sujet "Surfactant Interaction"

1

Alshaikh, Murtdha, Yeh Seng Lee et Berna Hascakir. « Anionic Surfactant and Heavy Oil Interaction during Surfactant-Steam Process ». Dans SPE Western Regional Meeting. Society of Petroleum Engineers, 2019. http://dx.doi.org/10.2118/195254-ms.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
2

Yang, C. Z. « Adjustment of Surfactant/Polymer Interaction in Surfactant/Polymer Flooding With Polyelectrolytes ». Dans SPE Enhanced Oil Recovery Symposium. Society of Petroleum Engineers, 1986. http://dx.doi.org/10.2118/14931-ms.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
3

« Wettability Alteration on Sandstone Reservoirs Containing Clay Minerals By The Addition Anionic Alkyl Ethoxy Carboxylate Surfactant ». Dans Indonesian Petroleum Association - 46th Annual Convention & Exhibition 2022. Indonesian Petroleum Association, 2022. http://dx.doi.org/10.29118/ipa22-e-298.

Texte intégral
Résumé :
With the increasing number of maturing fields, surfactant injection as an alternative EOR technique is essential to mobilize the remaining oil. The presence of clay in the sandstone causes the polar components in the oil to be adsorbed on the rock surface, causing a tendency towards wettability towards oil-wet. Therefore, the oil potential remaining in the reservoir, which is the target of surfactant injection, is more significant in oil reservoirs with oil-wet characteristics. For this reason, surfactants are needed that can alter the surface properties of oil-wet to water-wet and the ability of these surfactants to reduce IFT. The wettability alteration mechanism is one thing that needs to be understood more deeply, considering the critical function of surfactant injection in the EOR method. Studies on the effect of adding surfactants on sandstones containing clay minerals of certain types and concentrations need to be developed to assist the application of the oil recovery process at an advanced stage. Laboratory experiments are carried out by measuring the contact angle of sandstone samples conditioned according to reservoir characteristics with several concentrations of clay interacting with fluid. The measurement results show that the initial condition of the sandstone containing montmorillonite clay is water-wet, while the sandstone containing kaolinite clay is oil-wet. The presence of surfactant solutions gives a wettability alteration effect on sandstones containing montmorillonite clay to become more water-wet and sandstones containing kaolinite to be water-wet. Surfactant solution with a concentration of 1wt% gives a more significant wettability alteration effect than surfactants with the attention of 2wt%. The interaction of sandstones containing clay minerals with surfactant solution shows that the addition of surfactants can reduce the interfacial tension between oil and water. In contrast, surfactants with a concentration of 2wt% can reduce IFT greater than surfactants with the attention of 1wt%.
Styles APA, Harvard, Vancouver, ISO, etc.
4

Xu, Limin, Ming Han, Dongqing Cao et Alhasan Fuseni. « New Synergistic Surfactant Mixtures for Improving Oil Production in Carbonate Reservoirs ». Dans SPE Conference at Oman Petroleum & Energy Show. SPE, 2022. http://dx.doi.org/10.2118/200182-ms.

Texte intégral
Résumé :
Abstract The paper presents the development of new surfactant formulations composed of various low-cost and low-performance surfactants to make them high performance products for high temperature and high salinity carbonate reservoirs. The objective of this study is to optimize the surfactant chemistry by mixing different kinds of surfactants (ionic, nonionic, and amphoteric), which results in significant synergistic effects in interfacial properties to improve oil production at the given harsh conditions. The optimal mixing surfactant ratios were determined according to the brine-surfactant compatibility, microemulsion phase behavior, and the interfacial tension (IFT) between oil and surfactant solutions in high salinity brine and at 90˚C. Comprehensive performance of the surfactants was evaluated, including adsorption of the surfactants onto the carbonate rocks and the long-term stability at 95˚C. The coreflooding displacement experiments were performed using carbonate core plugs at 95˚C to evaluate the potential of the optimal mixing surfactants in improving oil production. Three formulations composed of two types of low-cost surfactants were developed in this study. The mixing surfactants were chosen based on moderate electrostatic interaction among the surfactants. It appeared the synergistic effect between the mixing surfactants was enhanced with increasing temperature. Although the IFT of the individual surfactants with crude oil was in the range of 100mN/m, a significant IFT reduction in the magnitude of 10−2 - 10−3 mN/m was observed by mixing the surfactants. A salinity scan showed that the IFT values maintained a value of 10−2 mN/m in a wide salinity range, which demonstrated the robustness of the surfactants mixtures. In microemulsion phase behavior studies, these mixed surfactant solutions in the presence of crude oil exhibited Winsor Type III emulsions. The static adsorptions of the mixed surfactants were lower than the individual surfactant adsorption. All this indicated the feasibility of these formulations for their applications in the harsh reservoir conditions. The results of coreflooding displacement tests demonstrated significant oil production improvement beyond water flooding. This work provides an efficient way to get surfactant formulations by mixing low-performance and low cost surfactants to obtain high performance in improving oil production under the harsh conditions.
Styles APA, Harvard, Vancouver, ISO, etc.
5

Salager, Jean-Louis, Jesus Ontiveros et Ronald Marquez. « How to use in practice a simplified HLDN linear equation for surfactant mixtures ». Dans 2022 AOCS Annual Meeting & Expo. American Oil Chemists' Society (AOCS), 2022. http://dx.doi.org/10.21748/uthm3166.

Texte intégral
Résumé :
In 1948 Winsor proposed the R ratio of the interactions at interface of the adsorbed surfactant molecules with oil (Aco) and water (Acw). At the R=Aco/Acw=1 condition, a three-phase behavior was attained. This occurrence in a continuous scan of a formulation variable, susceptible to change any interaction, was also mentioned to correspond to a low interfacial tension and low emulsion stability, among other properties. In 1975, Enhanced Oil Recovery initial studies indicated the necessity of attaining a very low interfacial tension minimum, i.e., the R=1 physicochemical balance situation presented by Winsor 25 years before. The experimental methods started expressing Winsor's optimum ratio condition (R=Aco/Acw=1) as the equivalent zero difference of interactions between the surfactant and oil and water, i.e., D=Aco-Acw=0. This motivated to numerically express the effects of formulation variables occurring in a petroleum reservoir to attain a minimum low IFT. The hydrophilic-lipophilic deviation (HLD=D/RT) from optimum formulation was expressed as a linear normalized equation as follows:HLDN = - ACN + Ks ln S €“ Kt T + C for ionic surfactants like sulfonates or carboxylatesHLDN = - ACN + K's S + K't T + C' for nonionic ethoxylated surfactants. Where S is the aqueous salinity, ACN is the oil nature (alkane carbon number), T is the temperature. The C and C' term includes the remaining variables as the head and tail characterization of the surfactant and eventual co-surfactant, the pressure, and the SOW composition. The C and C' also consist of the proportion X of the different species if there are two or more surfactants when a mixture is used. Herein, it is shown that there are different ways to express the effect of the mixture in the previous equations, that is, as a linear relation Kc X term over some range of mixture composition.
Styles APA, Harvard, Vancouver, ISO, etc.
6

Mehan, Sumit, Vinod K. Aswal et Joachim Kohlbrecher. « Structural study of surfactant-dependent interaction with protein ». Dans NANOFORUM 2014. AIP Publishing LLC, 2015. http://dx.doi.org/10.1063/1.4917622.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
7

Ray, D., et V. K. Aswal. « Tuning of depletion interaction in nanoparticle-surfactant systems ». Dans SOLID STATE PHYSICS : Proceedings of the 58th DAE Solid State Physics Symposium 2013. AIP Publishing LLC, 2014. http://dx.doi.org/10.1063/1.4872534.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
8

Feng, Lijie, et Xu Liang. « Implications of Shale Oil Compositions on Surfactant Efficacy for Wettability Alteration ». Dans SPE Middle East Unconventional Resources Conference and Exhibition. SPE, 2015. http://dx.doi.org/10.2118/spe-172974-ms.

Texte intégral
Résumé :
Abstract Surfactant selection is important for oil recovery in a hydraulically fractured reservoir. Two primary mechanisms, ion-pair coupling (cleaning) between surfactant and hydrocarbon and surfactant adsorption onto the rock surface (coating), were previously suggested to explain how surfactants can alter rock wettability, thus improving oil production. Because of the electrostatic interaction, acidic compounds in the oil tend to be adsorbed onto rock surface that is positively charged; whereas basic compounds are preferentially attracted to rock surface that is negatively charged. It has been discussed in previous literature that for wettability alteration for conventional formation rocks, the cleaning mechanism could be more efficient by inducing ion pairs between surfactants and oil compounds that have opposite charges, rather than the coating mechanism that mainly relies on how well surfactant covers the rock surface. In this research, 90 shale oils from various liquids shale plays, such as the Eagle Ford and the Wolfcamp, were tested for total acid number (TAN) and total base number (TBN). Cationic and anionic surfactants with low interfacial surface tension, along with Berea sandstone and Indiana limestone, were used to investigate the extent that TAN and TBN can be used as criteria to select surfactants. Oil recovery, imbibition, interfacial surface tension, and emulsion tendency were conducted to examine whether the cleaning mechanism holds true for shale oil saturated formation rocks. The results demonstrated that for carbonates with shale oil having a higher TAN, a cationic surfactant provided the potential to sweep more oil than an anionic surfactant. On the other hand, for sandstone with shale oil having a higher TBN, an anionic surfactant performed better than a cationic surfactant. Those observations appear to be consistent with the proposed cleaning mechanism and resonate with production data for thousands of wells from some major liquids-rich shale plays.
Styles APA, Harvard, Vancouver, ISO, etc.
9

Saputra, I. Wayan Rakananda, et David S. Schechter. « A Temperature Operating Window Concept for Application of Nonionic Surfactants for EOR in Unconventional Shale Reservoirs ». Dans SPE Annual Technical Conference and Exhibition. SPE, 2021. http://dx.doi.org/10.2118/206346-ms.

Texte intégral
Résumé :
Abstract Surfactant performance is a function of its hydrophobic tail, and hydrophilic head in combination with crude oil composition, brine salinity, rock composition, and reservoir temperature. Specifically, for nonionic surfactants, temperature is a dominant variable due to the nature of the ethylene oxide (EO) groups in the hydrophilic head known as the cloud point temperature. This study aims to highlight the existence of temperature operating window for nonionic surfactants to optimize oil recovery during EOR applications in unconventional reservoirs. Two nonylphenol (NP) ethoxylated nonionic surfactants with different EO head groups were investigated in this study. A medium and light grade crude oil were utilized for this study. Core plugs from a carbonate-rich outcrop and a quartz-rich outcrop were used for imbibition experiments. Interfacial tension and contact angle measurements were performed to investigate the effect of temperature on the surfactant interaction in an oil/brine and oil/brine/rock system respectively. Finally, a series of spontaneous imbibition experiments was performed on three temperatures selected based on the cloud point of each surfactant in order to construct a temperature operating window for each surfactant. Both nonionic surfactants were observed to improve oil recovery from the two oil-wet oil/rock system tested in this study. The improvement was observed on both final recovery and rate of spontaneous imbibition. However, it was observed that each nonionic surfactant has its optimum temperature operating window relative to the cloud point of that surfactant. For both nonionic surfactants tested in this study, this window begins from the cloud point of the surfactant up to 25°F above the cloud point. Below this operating window, the surfactant showed subpar performance in increasing oil recovery. This behavior is caused by the thermodynamic equilibrium of the surfactant at this temperature which drives the molecule to be more soluble in the aqueous-phase as opposed to partitioning at the interface. Above the operating window, surfactant performance was also inferior. Although for this condition, the behavior is caused by the preference of the surfactant molecule to be in the oleic-phase rather than the aqueous-phase. One important conclusion is the surfactant achieved its optimum performance when it positions itself on the oil/water interface, and this configuration is achieved when the temperature of the system is in the operating window mentioned above. Additionally, it was also observed that the 25°F operating window varies based on the characteristic of the crude oil. A surfactant study is generally performed on a single basin, with a single crude oil on a single reservoir temperature or even on a proxy model at room temperature. This study aims to highlight the importance of applying the correct reservoir temperature when investigating nonionic surfactant behavior. Furthermore, this study aims to introduce a temperature operating window concept for nonionic surfactants. This work demonstrates that there is not a "one size fits all" surfactant design.
Styles APA, Harvard, Vancouver, ISO, etc.
10

Bian, Yu, et Pinn-Tsong Chiang. « Effect of Hydrophobic/Hydrophilic Groups of Surfactants on Wax Deposition Studied by Model Waxy Oil System ». Dans SPE International Conference on Oilfield Chemistry. SPE, 2023. http://dx.doi.org/10.2118/213821-ms.

Texte intégral
Résumé :
Abstract It is well known that surfactants can be used as a wax dispersant, however, with properly adjusted ratios between the hydrophobic and hydrophilic groups of non-ionic surfactants, they can not only reduce the waxy oil pour point, but also reduce the detected wax appearance temperature (DWAT) and thus reduce wax deposition. Non-ionic surfactants with different numbers of hydrophobic/hydrophilic groups were studied as wax inhibitors using a model waxy oil system. Two model oils with different amounts and distribution of wax in dodecane were used in this study. High temperature gas chromatography (HTGC) was used to analyze the wax distributions. Surfactants with varying levels of ethoxylation and saturation were studied to find the most efficient structures for wax inhibition. A pour point tester was employed as an initial screening tool to determine the oil pour point and DWAT. A Turbiscan was used to evaluate the wax dispersing capabilities of the surfactants. Capillary flow through (CFT) wax deposition tests were then performed to verify the wax deposition reduction efficiency and to study the effect of the test parameters on wax deposition. The results showed that a surfactant with a moderate number of linear saturated alkane chains at an optimized dosage level can interact with the wax in dodecane, and thus reduce the pour point and DWAT. More alkane chains make the surfactants perform as a wax inhibitor (WI) but cause the solubility issues. Hydrophilic groups on the surfactants improve their solubility and interfere with the formation of wax crystals; however, having too many results in the surfactant self-assembling. The interaction between surfactants and wax changes with wax molecular weight (MW) and content. By optimizing the balance of hydrophobic and hydrophilic groups, the surfactant's wax inhibition performance can be improved. From this systematic study on the kinetic and dynamic behaviors of wax deposition, it was demonstrated that surfactants can be optimized to inhibit wax crystallization. By better understanding the relationship between their chemical structures and their performance, surfactant selection can be optimized with purpose-designed lab screening tests. Surfactants which are effective at wax inhibition could further mitigate wax deposition and keep the formulated WI package cost effective.
Styles APA, Harvard, Vancouver, ISO, etc.

Rapports d'organisations sur le sujet "Surfactant Interaction"

1

Gabitto, Jorge, et Kishore K. Mohanty. Surfactant-Polymer Interaction for Improved Oil Recovery. Office of Scientific and Technical Information (OSTI), janvier 2002. http://dx.doi.org/10.2172/789941.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
2

Unknown. SURFACTANT - POLYMER INTERACTION FOR IMPROVED OIL RECOVERY. Office of Scientific and Technical Information (OSTI), septembre 1997. http://dx.doi.org/10.2172/766780.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
3

Unknown. SURFACTANT - POLYMER INTERACTION FOR IMPROVED OIL RECOVERY. Office of Scientific and Technical Information (OSTI), octobre 1998. http://dx.doi.org/10.2172/766781.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
4

French, T. R., et C. B. Josephson. The effect of polymer-surfactant interaction on the rheological properties of surfactant enhanced alkaline flooding formulations. Office of Scientific and Technical Information (OSTI), février 1993. http://dx.doi.org/10.2172/10130748.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
5

P. Somasundaran. Mineral-Surfactant Interaction for Minimum Reagents Precipitation and Adsorption for Improved Oil Recovery. Office of Scientific and Technical Information (OSTI), septembre 2006. http://dx.doi.org/10.2172/902900.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
6

French, T. R., et C. B. Josephson. The effect of polymer-surfactant interaction on the rheological properties of surfactant enhanced alkaline flooding formulations. [Phase separation, precipitation and viscosity loss]. Office of Scientific and Technical Information (OSTI), février 1993. http://dx.doi.org/10.2172/6781205.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
7

Hirsa, Amir. Interaction of Surfactants with Shear Flows and Surface Waves. Fort Belvoir, VA : Defense Technical Information Center, septembre 1997. http://dx.doi.org/10.21236/ada628926.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
8

P. Somasundaran. MINERAL-SURFACTANT INTERACTIONS FOR MINIMUM REAGENTS PRECIPITATION AND ADSORPTION FOR IMPROVED OIL RECOVERY. Office of Scientific and Technical Information (OSTI), avril 2006. http://dx.doi.org/10.2172/882581.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
9

P. Somasundaran. MINERAL-SURFACTANT INTERACTIONS FOR MINIMUM REAGENTS PRECIPITATION AND ADSORPTION FOR IMPROVED OIL RECOVERY. Office of Scientific and Technical Information (OSTI), octobre 2004. http://dx.doi.org/10.2172/835274.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
10

P. Somasundaran. MINERAL-SURFACTANT INTERACTIONS FOR MINIMUM REAGENTS PRECIPITATION AND ADSORPTION FOR IMPROVED OIL RECOVERY. Office of Scientific and Technical Information (OSTI), avril 2005. http://dx.doi.org/10.2172/840105.

Texte intégral
Styles APA, Harvard, Vancouver, ISO, etc.
Nous offrons des réductions sur tous les plans premium pour les auteurs dont les œuvres sont incluses dans des sélections littéraires thématiques. Contactez-nous pour obtenir un code promo unique!

Vers la bibliographie