Fact-checked by Grok 2 weeks ago

Microemulsion

A microemulsion is a thermodynamically stable, isotropic liquid dispersion composed of , , , and often a cosurfactant, forming nanometer-sized droplets typically ranging from 10 to 100 in diameter, which results in optical and spontaneous formation without high-energy input. Unlike conventional emulsions, microemulsions exhibit low interfacial between the oil and water phases, enabling high solubilization capacity for both hydrophilic and hydrophobic substances. The term "microemulsion" was coined in 1959 by Jack H. Schulman and colleagues, building on earlier observations of stable oil-water mixtures stabilized by amphiphiles dating back to 1943. Microemulsions can adopt various microstructures, including oil-in-water (O/W), water-in-oil (W/O), and bicontinuous phases, where oil and water domains interpenetrate, depending on the composition ratios and the nature of the used. These systems are characterized by low viscosity, thermodynamic stability over time, and the ability to incorporate a wide range of active ingredients, making them distinct from macroemulsions that require mechanical agitation and are prone to . Microemulsions have found extensive applications in pharmaceuticals as vehicles, particularly for enhancing the of poorly soluble drugs through topical, oral, and parenteral routes, due to their ability to improve and protect payloads from . Beyond , they are utilized in for improved texture and stability, in the for emulsification, and in chemical processes like owing to their tunable interfacial properties. Recent advancements focus on formulating microemulsions with biocompatible components to address challenges in and regulatory approval for clinical use.

Fundamentals

Definition

A microemulsion is defined as a made of , , and that forms an isotropic and thermodynamically stable system with dispersed domain diameters ranging from approximately 10 to 100 nm. These systems consist of two immiscible liquids, typically and , mixed in comparable proportions to create low-viscosity, microheterogeneous fluids stabilized by amphiphilic that significantly reduce interfacial tension. The concept of such transparent dispersions was first described in 1943 by Hoar and Schulman, who observed the formation of clear water-in-oil systems upon addition of alcohols to emulsions. Microemulsions form spontaneously without requiring high-energy mechanical input, owing to the ultra-low interfacial tension (often near zero) enabled by the and optional cosurfactant components, which allows for easy into nanoscale structures. This thermodynamic stability distinguishes them from kinetically stable emulsions, as microemulsions maintain their properties indefinitely under given conditions without . The resulting domain sizes, smaller than the of (approximately 400–700 ), impart optical or translucency to the mixture. The typical composition includes an oil phase (hydrophobic solvent), an aqueous phase (water or ionic solution), surfactants (amphiphilic molecules that adsorb at the interface to minimize energy), and frequently a cosurfactant such as short-chain alcohols (e.g., n-butanol) to increase the interfacial flexibility and enlarge the stability region. Droplet diameters in these domains generally fall between 10 and 100 nm, enabling unique solubilization properties for both hydrophilic and lipophilic substances.

Properties

Microemulsions are characterized by ultralow interfacial between the oil and water phases, typically on the order of $10^{-3} mN/m or less, which facilitates the spontaneous formation of nanoscale domains without external energy input. This low , combined with the presence of and cosurfactants, enables high solubilization capacity for both hydrophobic and hydrophilic substances, allowing microemulsions to act as versatile carriers for poorly soluble compounds across a wide range of compositions. Physically, they exhibit Newtonian flow behavior, with viscosities generally ranging from 1 to 10 , akin to that of , due to their dilute, isotropic nature at low concentrations. Optically, microemulsions appear transparent or translucent because their droplet sizes, typically 10–100 nm, are smaller than one-twentieth of the visible wavelength (\lambda \approx 400–700 nm), preventing the observed in larger emulsions. Electrical conductivity in microemulsions varies significantly with their type: oil-in-water (o/w) systems display high conductivity (on the order of mS/cm) due to the continuous aqueous phase, while water-in-oil (w/o) systems show low conductivity (on the order of \muS/cm) dominated by the non-conducting oil phase; bicontinuous structures fall between these extremes. Microemulsions demonstrate thermodynamic over broad temperature ranges, often from -10°C to 100°C depending on composition, without or coalescence under conditions. This thermal resilience arises from the balanced intermolecular interactions that maintain the nanoscale structure across environmental variations.

Comparison with Other Systems

Microemulsions differ fundamentally from macroemulsions in their formation and . While microemulsions form spontaneously due to their thermodynamic , characterized by a negative of formation (ΔG < 0), macroemulsions require significant mechanical energy input for creation and are merely kinetically stable (ΔG > 0), leading to eventual creaming or over time. Additionally, microemulsion droplets typically range from 1 to 100 in diameter, rendering them transparent and low-viscosity, in contrast to the larger 0.1–100 μm droplets of macroemulsions, which appear opaque and milky. In comparison to nanoemulsions, microemulsions share nanoscale droplet sizes (typically <100 nm) but exhibit distinct stability profiles. Microemulsions achieve thermodynamic equilibrium without the need for high-energy processes, often relying on cosurfactants to lower interfacial tension, whereas nanoemulsions are kinetically stabilized and commonly produced via methods like high-pressure homogenization, potentially without cosurfactants, though they may destabilize over extended periods. This thermodynamic favorability (ΔG < 0) in microemulsions contrasts with the higher free energy state of nanoemulsions relative to phase-separated systems. Microemulsions also contrast with micelles and liposomes in their structural complexity and composition. Unlike micellar solutions, which consist of dilute surfactant aggregates forming simple spherical or cylindrical structures with limited oil incorporation and no tension gradient across the interface, microemulsions feature structured phases, including bicontinuous networks, that accommodate comparable volumes of oil and water through a duplex surfactant film with differing tensions on each side. Similarly, while liposomes are bilayer vesicles primarily composed of phospholipids for encapsulating hydrophilic or lipophilic drugs in an aqueous core, microemulsions employ surfactants and cosurfactants to create thermodynamically stable, isotropic dispersions without such vesicular architecture, enabling broader solubilization of both phases. A common misconception arises from the term "micro," which refers to the nanoscale domain sizes (1–100 nm) rather than visibility under a light microscope, as microemulsions appear clear and isotropic to the naked eye. Furthermore, unlike colloidal sols (solid particles dispersed in liquid) or gels (semi-solid networks), microemulsions remain fluid liquids with reversible phase behavior.

Components and Formation

Key Components

Microemulsions are formulated using four primary components: an oil phase, an aqueous phase, surfactants, and often cosurfactants, each selected to achieve thermodynamic stability and desired functionality. The oil phase typically consists of hydrocarbons such as decane or triglycerides like those found in vegetable oils (e.g., olive or sesame oil), or silicones, which provide hydrophobic domains for solubilizing lipophilic substances. Selection of the oil phase emphasizes compatibility with surfactants and the target application, as well as volatility for processes requiring evaporation. The aqueous phase is generally water or buffered solutions (e.g., phosphate-buffered saline at pH 7-9), forming hydrophilic regions for polar components. Its composition influences stability, with factors like pH and salinity playing key roles; for instance, high salinity concentrations favor water-in-oil microemulsions by altering ionic interactions at the interface. Surfactants, such as ionic types like , non-ionic ones like , or cationic variants, are essential for reducing interfacial tension between oil and water phases to near-zero values, enabling spontaneous formation. Selection criteria include biocompatibility, low toxicity, and the value, typically 4-6 for water-in-oil systems and 8-18 for oil-in-water systems. Cosurfactants, often short-chain alcohols like butanol or pentanol, are not always required but enhance the flexibility of the surfactant film at the interface, lower the packing ratio of amphiphiles, and broaden the phase boundaries for microemulsion existence. They are chosen for their amphiphilic nature and synergy with surfactants, typically used in amounts less than the surfactant to avoid destabilization. Typical compositions involve 10-40% oil and 10-40% water by weight, with the balance consisting of combined and , though these ratios are optimized via phase diagrams to ensure clarity and stability.

Formation Mechanisms

Microemulsions form through spontaneous emulsification upon mixing oil, water, , and often at ambient temperature, leading to self-organization that minimizes the system's free energy without requiring external mechanical shear or energy input. This process relies on the surfactants' ability to rapidly adsorb at the oil-water interface, creating a stabilized dispersion that achieves thermodynamic equilibrium. Unlike macroemulsions, which demand high-energy homogenization, microemulsions emerge as clear, isotropic fluids due to the nanoscale domain sizes (typically 10–100 nm) that prevent light scattering. The formation is predominantly entropy-driven, where the entropy gain from dispersing the immiscible phases outweighs the enthalpic penalty of mixing, facilitated by ultralow interfacial tensions (on the order of 10⁻³ to 10⁻⁵ mN/m). Surfactants form flexible interfacial films with high bending elasticity, allowing thermal fluctuations to promote disorder and stabilize structures such as spherical droplets or bicontinuous networks. This entropy dominance arises from the large interfacial area per surfactant molecule, enabling efficient packing and curvature without significant energy barriers. Cosurfactants, such as short-chain alcohols (e.g., n-butanol), play a crucial role by increasing the curvature elasticity of the surfactant film, which accommodates diverse geometries ranging from spherical micelles to saddle-like bicontinuous phases. They enhance molecular mixing at the interface, further reducing interfacial tension and promoting phase inversion, such as through changes in temperature (phase inversion temperature, PIT) or composition, where the optimal PIT for nonionic surfactants marks the transition from oil-in-water to water-in-oil structures. Experimentally, microemulsion formation is assessed by titrating components—typically adding water or oil dropwise to a surfactant-cosurfactant-oil mixture—until optical clarity indicates a single-phase system. Pseudo-ternary phase diagrams are constructed to identify optimal compositional ratios, plotting the boundaries of microemulsion regions against variables like surfactant concentration. Factors such as salinity influence curvature via the hydrophilic-lipophilic deviation () framework, where increasing salt concentration (e.g., NaCl) shifts the spontaneous curvature from positive (favoring oil-in-water) to negative (water-in-oil), optimizing formation at HLD ≈ 0.

Classification

Oil-in-Water

Oil-in-water (o/w) microemulsions feature discrete nanoscale oil droplets, typically ranging from 10 to 100 nm in diameter, dispersed throughout a continuous aqueous phase and stabilized by a monolayer of surfactant molecules adsorbed at the oil-water interface. These systems are characterized by a high water content, generally exceeding 50% by volume, which ensures the water phase dominates the overall composition and provides a hydrophilic external environment. The formation of such structures relies on achieving ultralow interfacial tension between the oil and water phases, often facilitated by the presence of surfactants and cosurfactants. Key characteristics of o/w microemulsions include high electrical conductivity, attributable to the percolation of the continuous water phase that allows ion mobility. They exhibit a hydrophilic external feel due to the surrounding aqueous medium, making them distinct from oil-continuous systems. Additionally, the dispersed oil droplets enable effective solubilization of lipophilic compounds within the aqueous matrix. Representative examples include systems formulated with non-ionic surfactants such as (polyoxyethylene(4) lauryl ether) in water containing soybean oil as the dispersed phase, which form stable o/w microemulsions suitable for applications like creams or oral formulations. Stability in o/w microemulsions is optimized with surfactants having a hydrophilic-lipophilic balance (HLB) value greater than 10, which favors the curvature required for oil droplet formation in water. These systems are particularly sensitive to temperature increases, which can alter surfactant hydrophilicity and induce phase inversion to water-in-oil (w/o) structures via mechanisms like the .

Water-in-Oil

Water-in-oil (w/o) microemulsions feature discrete nanodroplets of water, typically ranging from 10 to 100 nm in diameter, dispersed within a continuous oil matrix, where the oil phase constitutes more than 50% of the total volume. These nanodroplets are stabilized by a monolayer of surfactant molecules at the water-oil interface, which curves toward the water phase to accommodate the reverse micellar structure, enabling the solubilization of polar substances in a nonpolar environment. Key characteristics of w/o microemulsions include their thermodynamic stability, low electrical conductivity due to the insulating oil continuous phase (often in the range of microsiemens per centimeter), and inherently hydrophobic nature, which makes them suitable for hosting water-sensitive chemical reactions by confining aqueous reactants within isolated nanodroplets. This structure protects hydrophilic solutes from the surrounding oil, facilitating applications such as microreactors for nanoparticle synthesis or enzymatic processes that require an aqueous microenvironment. Representative examples of w/o microemulsions include systems formulated with (sorbitan monooleate) as the surfactant in isooctane as the oil phase, incorporating brine as the aqueous component, which forms stable dispersions used in extraction processes. Such systems are commonly employed in metal extraction, where heavy metals like copper or zinc achieve high extraction efficiencies through selective partitioning into the aqueous nanodroplets, followed by phase separation. In cosmetics, w/o microemulsions enhance the delivery of water-soluble actives, such as vitamins or peptides, into oil-based formulations for improved skin penetration and product stability. Stability in w/o microemulsions is primarily favored by surfactants with low hydrophilic-lipophile balance () values, typically below 8 (e.g., 3-6 for optimal curvature), which promote oil-continuous phases by enhancing lipophilicity. High salinity levels further stabilize these systems by screening electrostatic repulsions at the interface and altering the effective headgroup area of ionic surfactants, thereby reducing the spontaneous curvature toward the oil side. Phase inversion to or from w/o microemulsions can occur through , where changes in composition or conditions shift from (oil-in-water) to (water-in-oil) via intermediate bicontinuous states.

Bicontinuous

Bicontinuous microemulsions represent a distinct class within microemulsion systems, characterized by interpenetrating domains of oil and water that form infinite, interconnected channels rather than discrete droplets. The structure arises from saddle-shaped surfactant films that exhibit negative Gaussian curvature, enabling the creation of a sponge-like network where both oil and water phases percolate continuously throughout the system. These domains typically span 5-20 nm in size, allowing for a highly ordered yet fluid arrangement stabilized by the surfactant monolayer. This configuration is most prevalent at balanced oil-to-water volume ratios, approximately equal, which promotes the isometric nature of the phases without favoring one over the other. Key characteristics of bicontinuous microemulsions include moderate electrical conductivity, stemming from the percolation of both conductive (water) and non-conductive (oil) pathways, which results in a transitional regime between high-conductivity oil-in-water and low-conductivity water-in-oil systems. The interfaces maintain zero mean curvature on average, a feature that minimizes energetic penalties and supports the balanced curvature required for the networked topology. Additionally, these systems boast a high internal interfacial area, on the order of 100 m²/g, which dramatically enhances molecular interactions and facilitates rapid diffusion of solutes through the interconnected channels, enabling efficient transport across phases. Representative examples of bicontinuous microemulsions occur in middle-phase regions of ternary surfactant-oil-water systems, such as those formed with (AOT), water, and decane at optimal salinity levels around 0.6% NaCl, where the structure emerges as a single isotropic phase. These are often observed during transitions in phase diagrams, including the appearance of fish-eye phases or shifts from lamellar structures, highlighting the dynamic equilibrium in such formulations. The stability of bicontinuous microemulsions hinges on the flexibility of the surfactant films, which must accommodate the saddle-like deformations without fracturing, a property enhanced by surfactants with tunable bending rigidity. They are commonly associated with , where the middle phase coexists with excess oil and water, providing a thermodynamically stable environment under balanced formulation conditions.

Theoretical Foundations

Thermodynamic Stability

Microemulsions exhibit thermodynamic stability, meaning they form spontaneously and persist indefinitely without phase separation, as the Gibbs free energy change (ΔG) for their formation is negative. This stability arises from the balance in the equation ΔG = ΔH - TΔS, where ΔH represents the enthalpy change (primarily the positive interfacial energy cost), T is the absolute temperature, and ΔS is the entropy change (positive due to increased disorder from dispersing one phase into the other). In microemulsions, the entropy gain from the dispersed state dominates over the interfacial enthalpy penalty, ensuring ΔG < 0 and spontaneous formation under appropriate conditions. The free energy of mixing in microemulsions can be expressed as ΔG_mix = RT [x \ln x + (1-x) \ln (1-x)] + \gamma \Delta A, where R is the gas constant, T is temperature, x is the mole fraction of the dispersed phase, the first term captures the ideal entropy of mixing, and the second term accounts for the interfacial energy contribution with \gamma as interfacial tension and \Delta A as the increase in interfacial area. This spontaneity is enabled by ultra-low interfacial tension (\gamma < 0.1 , \mathrm{mN/m}), achieved through efficient surfactant packing at the interface, which minimizes the positive \gamma \Delta A term relative to the entropic contribution. Cosurfactants, such as medium-chain alcohols, further enhance stability by reducing \gamma and increasing configurational entropy through greater flexibility in the surfactant monolayer, allowing more dispersed configurations without energetic penalty. Unlike macroemulsions, which rely on kinetic stability against processes like creaming or coalescence, microemulsions achieve true equilibrium stability, rendering them reversible upon changes in composition or temperature without hysteresis. This equilibrium nature stems from the minimized free energy state, where the system resists separation due to the favorable balance of entropy and low interfacial energy.

Interfacial Tension and Packing

Microemulsions form due to the adsorption of surfactants at the oil-water interface, which dramatically lowers the interfacial tension (IFT) to ultralow values, typically ranging from 10^{-3} to 10^{-4} mN/m, enabling the creation of nanoscale droplets or bicontinuous structures with enormous interfacial areas on the order of 10^3 m^2 per gram of surfactant. This reduction occurs through the , which relates changes in IFT to surfactant accumulation at the interface; for systems following , the IFT can be modeled as \gamma = \gamma_0 - RT \Gamma \ln(1 + K c) where \gamma_0 is the IFT without surfactant, R is the gas constant, T is temperature, \Gamma is the maximum surface excess concentration, K is the adsorption equilibrium constant, and c is the bulk surfactant concentration. Such near-zero IFT minimizes the energy penalty for interface formation, promoting thermodynamic stability without requiring external energy input, unlike conventional emulsions. The molecular geometry of surfactants dictates the curvature and topology of microemulsion interfaces through the packing parameter P = v / (a l), where v is the volume of the hydrophobic tail, a is the effective cross-sectional area of the hydrophilic headgroup, and l is the extended length of the tail. Values of P < 1/3 correspond to conical-shaped surfactants that favor oil-in-water (o/w) microemulsions with spherical micelles exhibiting positive mean curvature; $1/3 < P < 1 indicates wedge-shaped molecules suitable for bicontinuous phases with saddle-like curvatures; and P > 1 promotes cylindrical or inverted cone geometries for water-in-oil (w/o) microemulsions with negative curvature. This parameter provides a geometric rationale for how surfactant molecular design influences aggregate morphology, with cosurfactants like short-chain alcohols often tuning P by altering headgroup hydration or tail volume. The flexibility of surfactant monolayers in microemulsions arises from their low bending rigidity, allowing spontaneous adoption of both positive and negative curvatures essential for diverse structures. This elasticity is quantified by the Helfrich bending energy functional, F_b = \int \left[ \frac{\kappa}{2} (2H - C_0)^2 + \bar{\kappa} K \right] dA, where \kappa is the bending modulus (typically 0.5–2 k_B T for microemulsions), \bar{\kappa} is the Gaussian modulus, H is the , C_0 is the spontaneous curvature (tunable by formulation), K is the , and the integral is over the surface. Low \kappa values facilitate the high interfacial fluctuations needed for microemulsion formation, contrasting with rigid bilayers in lamellar phases. The Winsor classification links these interfacial properties to phase behavior: Type I systems feature o/w microemulsions in equilibrium with excess oil, where the IFT between the microemulsion and oil phase is relatively higher (around 10^{-2} mN/m); Type III bicontinuous microemulsions coexist with both excess oil and water, exhibiting ultralow IFT (<10^{-3} mN/m) with both phases due to optimal packing and minimal bending energy; and Type II systems involve w/o microemulsions with excess water, mirroring Type I but with inverted curvature and IFT characteristics. This progression reflects how formulation variables like salinity or temperature shift C_0 and IFT, transitioning between types while maintaining near-zero effective tension for stability.

Phase Behavior

The phase behavior of microemulsions is characterized by reversible transitions between distinct structures, influenced primarily by temperature, salinity, and composition, which alter the interfacial curvature and stability of the dispersed phases. Temperature plays a pivotal role in nonionic surfactant-based microemulsions, where increasing temperature shifts the hydrophilic-lipophilic balance toward oil solubility, leading to a phase inversion from oil-in-water (o/w) to water-in-oil (w/o) structures. This transition occurs at the phase inversion temperature (PIT), a characteristic point where the spontaneous curvature of the surfactant film changes sign due to dehydration of the hydrophilic headgroups, enabling efficient emulsification when the system is heated through the PIT and then cooled. For nonionic surfactants like polyethoxylated alcohols (e.g., C12E4), the PIT correlates closely with the cloud point, above which the surfactant becomes less hydrophilic, promoting w/o microemulsions; typical PIT values range from 20–80°C depending on the ethoxylation degree. Salinity effects are prominent in ionic surfactant systems, where added electrolytes screen the electrostatic repulsion between charged headgroups, reducing the effective headgroup area and favoring negative curvature for w/o microemulsions. As salinity increases, the system progresses from (o/w microemulsion coexisting with excess oil) through a three-phase region (, with a bicontinuous microemulsion) to (w/o microemulsion coexisting with excess water), with the optimal salinity marking the point of minimal interfacial tension in the middle-phase regime. Critical salinity for phase splitting occurs when the solubilization capacity is exceeded, leading to macroscopic separation; for anionic surfactants like , this threshold can be around 5–10 wt% NaCl, depending on cosurfactant presence. Composition sweeps, particularly varying surfactant concentration at fixed oil-to-water ratios, reveal characteristic "fish" diagrams that delineate the one-phase microemulsion channel flanked by two- and three-phase regions. In these diagrams, the "fish tail" at low surfactant concentrations shows a narrow channel of isotropic one-phase microemulsions, widening into a three-phase body as surfactant increases, before narrowing again at high concentrations where ordered phases dominate; this behavior arises from the balance of curvature and packing constraints. For systems like water/n-dodecane/, the one-phase region spans surfactant fractions of 5–20 wt% at optimal temperature or salinity. The packing parameter briefly influences these sweeps by dictating curvature preferences, but observable transitions reflect macroscopic equilibrium. Microemulsions often coexist with excess oil or water phases outside the one-phase channel, forming macroscopically separated layers while maintaining ultralow interfacial tensions at the boundaries. At high surfactant concentrations (typically >20 wt%), viscoelastic transitions to lamellar phases occur, driven by increased bending rigidity and reduction, resulting in stacked bilayers that separate oil and water domains; this is evident in systems like //, where the lamellar phase stabilizes above the microemulsion channel.

Phase Diagrams

Construction Methods

Phase diagrams for microemulsions are constructed using a combination of experimental and computational approaches to map the compositional regions where stable, isotropic microemulsion phases form. For systems involving , , and , diagrams are plotted in equilateral triangular coordinates, with each vertex representing 100% of one component. The microemulsion region is identified as the area of low and optical , often encompassing oil-in-water, water-in-oil, or bicontinuous structures. Experimental construction of diagrams typically employs methods, where fixed ratios of and oil are progressively diluted with (or ) until the mixture achieves clarity, indicating the boundary. Visual inspection confirms the isotropic nature of the microemulsion by observing against a source, while at 10,000–20,000 rpm for 30–60 minutes accelerates along tie-lines, allowing separation of coexisting phases for further analysis. measurements, using a probe to scan along dilution lines, distinguish phase types: high conductivity (mS/cm) signals oil-in-water microemulsions due to ionic pathways in the aqueous , while low values (µS/cm) indicate water-in-oil structures; bicontinuous regions show intermediate, percolating conductivities. scans, monitored via at 600 nm, complement these by detecting cloudiness at boundaries where macroscopic separation occurs. When cosurfactants such as short-chain alcohols are included, pseudo-ternary diagrams simplify the quaternary system by fixing the cosurfactant-to- weight (e.g., 1:1 or 2:1) and treating the mixture as a single "effective " vertex. These diagrams are constructed similarly via , revealing expanded microemulsion regions due to reduced interfacial tension, with the largest isotropic areas often at optimal ratios determined iteratively. For full systems (, cosurfactant, oil, ), tetrahedral diagrams are challenging to visualize, so slice diagrams are used, projecting fixed-ratio cuts (e.g., constant cosurfactant concentration) onto planes to map phase behavior. Computational modeling aids this, with software like simulating phase equilibria through phase-field methods that solve coupled diffusion equations for component concentrations, predicting boundaries without exhaustive experiments. Fish-tail diagrams, particularly for nonionic , are generated by scanning and (or surfactant concentration) at fixed oil-to-water ratios (e.g., 1:1). These plot the lower of the single-phase microemulsion , forming a "fish tail" shape where the phase widens at optimal conditions and narrows with increasing or , reflecting transitions; the tail tip marks the minimum concentration for stability. Anionic yield narrower tails, while nonionics show broader, -sensitive regions.

Interpretation

Phase diagrams for microemulsions are graphical representations that map the equilibrium phase behavior of surfactant-oil-water systems, typically as functions of composition, temperature, or salinity, allowing prediction of stable microemulsion regions and phase transitions. The one-phase region, often shaded in these diagrams, denotes the area of thermodynamic stability where a single, isotropic microemulsion phase forms without phase separation, reflecting the system's ability to solubilize oil and water at ultralow interfacial tensions on the order of 10^{-4} mJ/m². The width of this region indicates surfactant efficiency, defined as the minimum surfactant concentration (φ_s^*) required to achieve solubilization of equal volumes of oil and water; for instance, non-ionic surfactants like C_{12}E_5 (pentaethylene glycol monododecyl ether) typically yield broader one-phase regions due to their temperature-tunable hydrophile-lipophile balance (HLB), whereas ionic surfactants often require co-surfactants or electrolytes to expand the region, resulting in narrower stability domains under similar conditions. In multiphase regions of the diagram, tie-lines connect the compositions of coexisting phases, such as a microemulsion in with excess or , enabling of phase compositions and relative volumes via the : the position of a bulk composition along a tie-line determines the volume fractions of each , with the distance to each end inversely proportional to the phase amounts. This interpretation is central to classifying Winsor types of —Winsor I (Type I), where an oil-in- (o/w) microemulsion coexists with excess ; Winsor II (Type II), featuring a -in- (w/o) microemulsion with excess ; and Winsor III (Type III), characterized by a bicontinuous microemulsion in with both excess and , forming a three- tie-triangle with phase compositions at the vertices. These types arise from curvature transitions at the - , driven by formulation variables, and the diagrams predict transitions between them, such as from o/w to bicontinuous as conditions favor zero . The "fish diagram," a common binary cut of the phase prism at equal oil-to-water ratios (φ_w = φ_o), provides a simplified view for identifying optimal formulations, where the midline or "" at the balance temperature (T_0) corresponds to ultra-low interfacial tension and maximum solubilization capacity, ideal for applications requiring efficient oil recovery or . Along this diagram, temperature dependence reveals inversion loci: for non-ionic , increasing temperature shifts from Winsor I (negative , o/w) through Winsor III (zero , bicontinuous) to Winsor II (positive , w/o), with the three-phase body forming the "fish tail" at lower concentrations; this tunability stems from the dehydration of headgroups, altering spontaneous . In contrast, ionic systems often use as the tuning parameter, with similar inversion patterns but potentially sharper transitions. Recent advances as of include the integration of to predict phase boundaries from limited experimental data and advanced in situ techniques like under confinement to study nanoscale phase behavior in porous media. Despite their utility, phase diagrams have limitations in interpretation, as they depict states and ignore kinetic factors such as times or metastable states that may persist outside the predicted regions. Additionally, most diagrams are 2D projections of higher-dimensional phase prisms (e.g., ternary composition with or ), simplifying complex 4D behaviors and potentially overlooking nuances in cosurfactant effects or multicomponent interactions. These tools thus excel in predicting stable microemulsion formation but require complementary experimental validation for dynamic processes.

Applications

Industrial Uses

Microemulsions play a significant role in (EOR) by forming low-interfacial-tension systems that mobilize residual oil trapped in reservoir pores. These formulations, often involving like combined with cosurfactants such as n-butanol, reduce oil-water interfacial tension to ultralow levels (10⁻²–10⁻³ mN/m), enabling efficient displacement through flooding. For instance, in low-permeability reservoirs, microemulsion systems with internal olefin sulfonates have achieved additional oil of up to 30% over conventional water flooding, as demonstrated in laboratory core flooding experiments and field trials like those at the Jidong oilfield in , where cumulative oil production increased by 3000 tons. In and , microemulsions enable the creation of transparent formulations such as shampoos and creams that solubilize fragrances and oils without . Their nanoscale droplet sizes (10–100 nm) contribute to optical clarity and thermodynamic , while nonionic like polyoxyethylene (20) oleyl ensure low irritation potential for and applications. Examples include clear perfuming gels incorporating 33% oil with for enhanced spreadability and long-term exceeding one year at , and conditioners using amino-functional silicones for improved conditioning without residue. The utilizes food-grade microemulsions as emulsifiers to incorporate flavors and oils into aqueous-based products like beverages, achieving clear, stable dispersions. Lecithin-based systems, for example, solubilize hydrophobic bioactives such as essential oils, enhancing their while maintaining low and suitable for commercial drinks. These formulations leverage the isotropic nature of microemulsions to prevent creaming or , supporting applications in flavor delivery without altering sensory attributes. Microemulsions enhance cleaning products, including detergents, by providing superior grease solubilization through their oil-in-water structures stabilized by surfactants like Pluronic block copolymers. In formulations with essential oils such as or , they effectively remove oily soils, reducing carbon content by up to 52.9% in cleaning tests on metal surfaces, while also serving in agrochemicals for controlled release via nanoscale encapsulation. This dual action of emulsification and dispersion improves efficacy in both household and industrial cleaners. In paints and coatings, microemulsions facilitate better pigment dispersion and enable low-volatile organic compound (VOC) formulations by stabilizing nanoscale domains of resins and solvents. Systems polymerized with surfactants like cetyltrimethylammonium bromide produce particles as small as 0.02–0.14 µm, leading to homogeneous, high-intensity colors and reduced environmental impact compared to traditional solvent-based paints. Acrylate lattices derived from such microemulsions exemplify their role in durable, eco-friendly coatings.

Pharmaceutical Applications

Microemulsions have emerged as versatile carriers in pharmaceutical formulations, particularly for enhancing the solubilization of poorly water-soluble drugs, which often belong to (BCS) classes II and IV. Their large interfacial area and low interfacial tension enable the incorporation of lipophilic actives into the oil phase or at the , significantly improving aqueous and compared to conventional s or suspensions. For instance, cyclosporine, a hydrophobic immunosuppressant, is solubilized in microemulsion-based formulations like Neoral, where it achieves effective oral absorption by forming a fine upon dilution in gastrointestinal fluids. In systems, microemulsions facilitate various administration routes. For topical applications, oil-in-water (o/w) microemulsions promote penetration of actives like ibuprofen or by reducing the 's barrier resistance and enhancing partitioning into the . Oral delivery benefits from self-microemulsifying systems (SMEDDS), which spontaneously form microemulsions in the gut, promoting lymphatic and bypassing first-pass ; examples include formulations of and for antiretroviral therapy. Parenteral routes utilize stable microemulsions for intravenous administration, such as those loading at concentrations up to 2.5 mg/mL, offering Cremophor EL-based systems with improved tolerability over traditional solvents. Additionally, (α-) microemulsions serve as carriers, enhancing mucosal delivery and stability in oral or topical contexts. Key advantages of microemulsions in pharmaceuticals include controlled release profiles, which sustain levels and reduce dosing , and decreased through lower concentrations compared to macroemulsions. These systems also mitigate issues like , ensuring consistent . However, challenges such as potential -induced , particularly from ionic types, can occur; this is often addressed by employing non-ionic like polysorbates, which exhibit better and lower . Regulatory approval underscores their clinical viability, with FDA-cleared products like Neoral (cyclosporine microemulsion) demonstrating enhanced over predecessors like Sandimmune, achieving up to 50% higher in transplant patients.

Emerging Developments

Recent advancements in microemulsion technology have integrated to enhance systems, particularly for cancer therapies. Hybrid microemulsions incorporating gold nanoparticles (AuNPs) in oil-in-water (o/w) configurations have shown promise for simultaneous imaging and treatment, where AuNPs serve as contrast agents in (MRI) while enabling photothermal ablation of tumor cells. Studies have demonstrated improved tumor accumulation via the and higher cellular uptake compared to free AuNPs, with minimal toxicity to healthy tissues. Sustainable formulations of microemulsions are gaining traction through the use of bio-based surfactants derived from renewable sources like sugars, aiming to reduce environmental impact in applications such as enhanced oil recovery (EOR). These green surfactants, such as alkyl polyglucosides, lower interfacial tension to below 10^{-2} mN/m, facilitating oil mobilization while being biodegradable and non-toxic. In 2023 research, bio-based surfactant microemulsions achieved 15-20% additional oil recovery in core flooding experiments under reservoir conditions, outperforming synthetic counterparts in salinity tolerance. Enzyme encapsulation within microemulsions has emerged as a key strategy for biocatalysis, protecting enzymes from denaturation and enabling efficient substrate conversion in non-aqueous media. Reverse microemulsions have been used to immobilize lipases, maintaining over 90% activity after multiple cycles in biodiesel production. A 2023 report highlighted enzyme-loaded microemulsions that catalyzed esterification reactions with 95% yield, demonstrating reusability up to 10 times without significant loss in performance. In and delivery, bicontinuous microemulsions provide a protective matrix for sensitive nucleic acids like siRNA, shielding them from degradation while facilitating endosomal escape. These structures, with interconnected water and oil domains, enhance siRNA loading and achieve targeted silencing in hepatocytes. For vaccines, emulsion-based adjuvants have been explored to boost immune responses, with squalene-in-water systems eliciting higher titers in preclinical models compared to non-adjuvanted formulations. Microemulsion-based approaches are advancing , particularly through soil washing to extract hydrophobic pollutants like hydrocarbons. Biodiesel-derived microemulsions reduce interfacial tension to ultralow levels (<10^{-3} mN/m), desorbing up to 70% of from contaminated soils in batch tests. Additionally, microemulsions enhance CO2 solubilization for carbon capture, with surfactant-stabilized systems increasing CO2 by 25-30% in aqueous phases under mild pressures, aiding post-combustion capture processes. Despite these innovations, challenges in persist, as transitioning from lab-scale to often results in instability and inconsistent droplet sizes due to mixing inefficiencies. As of 2025, AI-optimized diagrams have addressed this by predicting optimal ratios with high accuracy, reducing formulation trials and enabling custom designs for specific applications. models trained on hydrophile-lipophile balance data have successfully mapped microemulsion regions, forecasting stability under varying temperatures and salinities.

History

Early Discoveries

In 1943, researchers T. P. Hoar and J. H. Schulman at the observed clear, stable mixtures of oil, water, and , which formed spontaneously upon mixing without requiring high-shear emulsification. These transparent dispersions exhibited low interfacial tension at the oil-water interface, allowing nanoscale domains to persist indefinitely, marking the first documented recognition of such thermodynamically stable systems, though the term "microemulsion" was not yet applied. During the 1950s, J. W. McBain introduced the concept of "swollen micelles" to characterize similar oil-swollen aggregates in aqueous solutions, emphasizing their role in solubilizing hydrocarbons through expanded micellar structures. Early investigations by McBain and collaborators also examined films and low-tension interfaces, revealing how reduced interfacial energy to facilitate these isotropic, low-viscosity phases without external energy input. In the , L. M. and colleagues advanced understanding through studies of phase behavior in systems, mapping regions where water, , and formed stable, isotropic liquids via electron microscopy and phase diagrams. This work highlighted spontaneous structuring in nonionic and anionic mixtures, paving the way for practical insights. In 1968, W. R. Foster proposed using these low-tension systems for , demonstrating their potential to mobilize residual by achieving ultralow interfacial tensions on the order of 10^{-3} mN/m. Key experiments in this era relied on conductivity and viscosity measurements to confirm the nanoscale domains in these mixtures; for instance, high electrical in oil-continuous systems indicated percolating water channels, while Newtonian profiles underscored the absence of large droplets, distinguishing them from macroemulsions that require intense mechanical for formation. Thermodynamic favorability of these systems, later formalized, stemmed from the entropy-driven mixing enabled by .

Terminology Evolution

The term "microemulsion" was first introduced in the late by J.H. Schulman and colleagues to describe transparent, thermodynamically stable dispersions of and in the presence of and cosurfactants, distinguishing them from kinetically stable macroemulsions that scatter light and appear opaque. This terminology emerged from early studies on low-interfacial-tension systems, where Schulman et al. noted the spontaneous formation of these isotropic solutions without high-energy input, unlike conventional emulsions. By the early , researchers including I. Danielsson and P. Stenius further characterized these systems through colloidal and interfacial analyses, emphasizing their nanoscale domain sizes (typically 1–100 nm) and optical clarity. In the 1970s, P.A. Winsor's earlier classification framework from 1948—categorizing systems into Types I (oil-in-water microemulsion in with excess oil), II (water-in-oil microemulsion with excess water), and III (bicontinuous middle-phase microemulsion coexisting with both excess oil and water)—gained widespread adoption for microemulsions, providing a basis for understanding phase behavior and stability. This classification helped standardize discussions at emerging international forums, highlighting the reversible phase transitions driven by variables like , , or cosurfactant concentration. In 1976, L.E. Scriven introduced the concept of "bicontinuous," describing interconnected oil and water domains in Type III systems without discrete droplets, a concept formalized through thermodynamic modeling of minimal surfaces and later validated by techniques. Debates arose over , particularly distinguishing microemulsions from "nanoemulsions," which are kinetically stabilized fine dispersions formed by high-energy methods, lacking thermodynamic stability; these discussions intensified in the as applications in prompted clearer terminological boundaries. From the 1990s onward, the Hydrophilic-Lipophilic Difference (HLD) framework developed by J.L. Salager in the late 1970s and refined into the HLD-Net Average Curvature (NAC) model in the 2000s provided a quantitative tool for predicting and standardizing microemulsion formulation by balancing surfactant hydrophilicity and against oil type and salinity. This approach enabled optimal Winsor III conditions for applications like and extraction processes. The International Union of Pure and Applied Chemistry (IUPAC) formalized the definition in the 1990s as "a made of water, oil, and that is an isotropic and thermodynamically stable system with dispersed domain diameter generally in the 1–100 nm range," underscoring stability over kinetic aspects. Terminological controversies persisted, notably confusion with "miniemulsions" in polymerization contexts, where miniemulsions refer to metastable nano-sized droplets (20–500 nm) stabilized by surfactants and hydrophobes for controlled radical polymerization, unlike the equilibrium structures of microemulsions. Current consensus, as reflected in recent reviews, avoids interpreting "micro" as implying microscopic visibility, instead emphasizing the nanoscale, non-scattering nature and thermodynamic favorability to prevent conflation with larger, unstable emulsions.

References

  1. [1]
  2. [2]
  3. [3]
    Exploring the Versatility of Microemulsions in Cutaneous Drug ... - NIH
    May 21, 2023 · Microemulsions are novel drug delivery systems that have garnered significant attention in the pharmaceutical research field.3.1. Surfactants · 3.2. Co-Surfactants · 5. Ternary Phase Diagrams Of...
  4. [4]
  5. [5]
    Nano- and Microemulsions in Biomedicine: From Theory to Practice
    The microemulsion is the thermodynamically stable phase from oil, water, surfactant and co-surfactant which forms the surface of drops with very small surface ...
  6. [6]
  7. [7]
    Correlation between interfacial tension and microemulsion structure ...
    At equal solubilities in both phases, the interfacial tension tends to a minimum of ∼ 0.001 mN m− 1 which is a condition for forming of a three-phase (Winsor- ...
  8. [8]
    Water solubilization capacity, interfacial compositions and ...
    The water solubilization capacity of a particular microemulsion depends upon the partitioning of cosurfactants among water, oil and interfaces, chain length ...
  9. [9]
    Microemulsion Rheology: Newtonian and Non-Newtonian Regimes
    Oct 4, 1981 · At a fixed shear rate or shear stress, microemulsion viscosity often goes through two maxima as the salinity of the mixing brine increases and ...
  10. [10]
    Development of an osmoprotective microemulsion as a therapeutic ...
    Jul 25, 2022 · We now know that microemulsions are transparent, thermodynamically stable nanosystems with small droplet sizes ... Tyndall effect and high ...
  11. [11]
    Microemulsion Microstructure(s): A Tutorial Review - PMC - NIH
    Microemulsions are thermodynamically stable, transparent, isotropic single-phase mixtures of two immiscible liquids stabilized by surfactants (and possibly ...
  12. [12]
    Thermal stability control of the water-in-diesel microemulsion fuel ...
    The present study discusses how to control the thermal stability of the microemulsion fuels in a temperature range from −21 °С to +73 °С.
  13. [13]
    Introductory Chapter: Microemulsions - IntechOpen
    May 24, 2017 · Simple electrical conductivity (EC) measurements can be used to identify whether the oil/water or both phases are continuous [5, 6]. The EC of ...Missing: ww/ | Show results with:ww/<|separator|>
  14. [14]
    None
    ### Summary of Microemulsions from https://www.chm.bris.ac.uk/eastoe/Surf_Chem/3%20Microemulsions.pdf
  15. [15]
    Nanoemulsions versus microemulsions: terminology, differences ...
    Dec 22, 2011 · Microemulsions have a lower free energy than the phase separated state, whereas nanoemulsions have a higher free energy. The two states are ...
  16. [16]
    Microemulsions versus micelles - ScienceDirect.com
    Microemulsions have a duplex film with different tensions at each side, while micelles have a limit in the ratio of alcohol to soap and amount of oil.
  17. [17]
    Liposomes, micelles and microemulsions as new delivery systems ...
    This review describes the design and performance of specialized delivery systems, such as liposomes, micellar solutions and microemulsions, ...
  18. [18]
    Full article: Role of microemuslsions in advanced drug delivery
    Microemulsions are isotropic and transparent systems, thermodynamically stable, and are comprised of oil, water, and a surfactant, mostly in association with a ...Missing: seminal | Show results with:seminal
  19. [19]
    None
    ### Extracted Examples of Oil Phases in Microemulsions
  20. [20]
    RJPT - Role of Surfactant and Co-surfactant in Microemulsion: A Review
    ### Summary of Roles of Cosurfactants in Microemulsions
  21. [21]
    Microemulsion Delivery Systems with Low Surfactant Concentrations
    In our research, such a composition was shown by MEs with the ratio of oil/surfactants + cosurfactant equal to 30:70 and 40:60 wt % (marked as points in the ...
  22. [22]
  23. [23]
  24. [24]
  25. [25]
    Oil-in-water microemulsions from association structures of surfactant ...
    The present review discusses new developments and optimization of demulsification processes in oil and gas applications, and highlights the critical parameters.
  26. [26]
    Preparation and Characterization of Microemulsions Based on ... - NIH
    Sep 25, 2020 · After screening the formula, we researched several characteristics of the prepared oil-in-water microemulsion, including electrical conductivity ...
  27. [27]
    Emulsions: making oil and water mix – AOCS
    Sep 27, 2024 · The HLB scale ranges from 0 to 20, with 10 corresponding to an emulsifier that is equally attracted to water and oil. Emulsifiers with HLB ...
  28. [28]
    [PDF] Design of a Microemulsion-based Drug Delivery System for ...
    Hydrophilic drugs solubilize mainly in water phase in the water-in-oil microemulsion droplets and hydrophobic drugs in oil droplets of oil-in-water.
  29. [29]
    The effect of sodium chloride on the formation of W/O ...
    By a simple control of surfactant concentration, the non-ionic microemulsions Water/polyoxietilen(4)laurylether (Brij 30)/Ethyl acetate (EtOAc) system can ...
  30. [30]
    Biocompatible microemulsions and their prospective uses in drug ...
    Sep 21, 2007 · ... Brij-30, and Brij-92) and cosurfactants (ethanol and isopropanol ... microemulsion with soy oil as the oil phase and DSPE–PEG as surfactant.
  31. [31]
    Development of oil-in-water microemulsions for the oral delivery of ...
    Oct 1, 2013 · Generally, surfactants with HLB values between 8 and 20 are able to form O/W MEs, while W/O MEs are formed when the HLB range is 4–7 (Lawrence ...
  32. [32]
    [PDF] MICROEMULSION - RJPN
    Microemulsion may be defined as isotopically clear, thermodynamically stable dispersion of two immiscible liquids (oil and water) which are stabilized by ...<|control11|><|separator|>
  33. [33]
    [PDF] Using the dynamic Phase Inversion Temperature (PIT) as a ... - HAL
    Dec 21, 2021 · Increasing the temperature leads to a decrease of the droplet size due to a decrease of the O/W interfacial tension. At the PIT, the SOW system.<|control11|><|separator|>
  34. [34]
  35. [35]
  36. [36]
    Effects of oil on the curvature elastic properties of nonionic surfactant ...
    Apr 11, 2006 · It adopts a bicontinuous structure where the midplane of the surfactant film forms a three dimensional simply connected surface of average zero ...<|control11|><|separator|>
  37. [37]
    Review on Some Confusion Produced by the Bicontinuous ...
    Mar 2, 2023 · Microemulsion refers to a single-phase system that does not contain distinct oil or water droplets but at least four different structures with ...
  38. [38]
    Microemulsions as model systems - ScienceDirect.com
    A common approach has been to describe a microemulsion in terms of a flexible surfactant film separating aqueous and oil domains. In the Helfrich ...
  39. [39]
    Percolation in an Oil-Continuous Microemulsion | Phys. Rev. Lett.
    Oct 28, 1985 · Conductivity data are in agreement with a dynamic percolation picture and inconsistent with the static percolation expected from a bicontinuous ...
  40. [40]
    Bicontinuous structure zones in microemulsions - Nature
    Oct 22, 1981 · ... conductivity variations with water content increasing in certain w/o microemulsions can be accounted for on the basis of the Percolation and ...
  41. [41]
    [PDF] in microemulsions - Société Chimique de France
    In addition, they possess a large internal oil-water interfacial area, on the order of 100 m²/g, at which reaction between oil and water soluble reagents can ...
  42. [42]
    Structural evolution of bicontinuous microemulsions
    Sep 1, 1991 · Structural evolution of bicontinuous microemulsions. Click to ... The Journal of Chemical Physics 2001, 115 (21) , 10036-10044. https ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  43. [43]
    Microemulsion | KRÜSS Scientific
    Unlike emulsions, they form spontaneously or with only a small amount of energy input, and in any case without intensive shaking or stirring.
  44. [44]
    [PDF] Insights of Microemulsions – A Thermodynamic Comprehension
    Nov 10, 2016 · The change in the Gibbs free energy of the system (ΔG) in terms of enthalpy and entropy can be written thermodynamically as: ΔG = ΔH− T ΔS. (15) ...Missing: dominates | Show results with:dominates
  45. [45]
    Microemulsion and L3 Phases of Water−Oil−Surfactant Mixtures
    The solid lines depict the coexistence lines of the isotropic phases L3 and microemulsion (μE) with water (W) and oil (O). The broken lines circumscribe the ...
  46. [46]
    The Correlation between Phase Inversion Temperature In Emulsion ...
    Phase inversions observed in thermoresponsive pickering emulsions stabilized by surface functionalized colloidal silica.
  47. [47]
    [PDF] SOME PHYSICOCHEMICAL ASPECTS OF MICROEMULSION ...
    ASPECTS OF MICROEMULSION FLOOING. 415. D. Optimal Salinity. Three optimal salinities have been so far defined, Cm (see IV D3), Cγ and Cφ. For the system 3% 63 ...
  48. [48]
    Effect of Pressure and Temperature on the Phase Behavior of ...
    Nov 1, 1990 · Temperature stability was ±0.1°C. ... Besides the compressibility effect, however, the physical properties of microemulsions are functions of the ...
  49. [49]
  50. [50]
  51. [51]
    Advance of Microemulsion and Application for Enhanced Oil Recovery
    This study initially delineates the fundamental concepts, classifications, formation mechanisms, advantages, and preparation methodologies of microemulsions.
  52. [52]
  53. [53]
    [PDF] Number 7 Microemulsions in Cosmetics - IFSCC MONOGRAPH
    Tyndall effect. Rayleigh derived the following equation for the ... tion at the oil droplet surface and by decreasing the droplet size and, thus ...
  54. [54]
    Microemulsions: A Potential Delivery System for Bioactives in Food
    Microemulsions are thermodynamically stable, transparent, low viscosity, and isotropic dispersions consisting of oil and water stabilized by an interfacial ...
  55. [55]
    Novel Microemulsions with Essential Oils for Environmentally ... - NIH
    Aug 26, 2023 · In this work, a novel microemulsion was prepared as a solution to replace the commonly used organic solvent mixtures for cleaning grime and oily ...
  56. [56]
    (PDF) Uses and applications of microemulsions - ResearchGate
    Aug 5, 2025 · Emulsions and microemulsions are presented in this review article, in which we give the differences between these two mixtures and we give the ...
  57. [57]
    [PDF] neoral - accessdata.fda.gov
    Neoral is an oral formulation of cyclosporine that immediately forms a microemulsion in an aqueous environment. Cyclosporine, the active principle in Neoral, ...
  58. [58]
    Microemulsions: Unique Properties, Pharmacological Applications ...
    Nov 11, 2021 · This review highlights the properties and several recent developments in the use of microemulsions for medical treatment purposes including targeted drug ...
  59. [59]
    Microemulsion co-delivering vitamin A and vitamin E as a new ...
    Jan 7, 2020 · We developed a water-in-oil microemulsion containing vitamin A (retinol) and vitamin E (α-tocopherol), which serves as a multifunctional nanosystem that co- ...
  60. [60]
    A Novel Paclitaxel Microemulsion Containing a Reduced Amount of ...
    The purpose of this study was to prepare a novel paclitaxel (PTX) microemulsion containing a reduced amount of Cremophor EL (CrEL) which had similar ...
  61. [61]
    Topical Microemulsions: Skin Irritation Potential and Anti ... - NIH
    Nonionic surfactants, frequently utilized in skincare products, have a lower likelihood of causing irritation [91]. They are also formulated in combination ...
  62. [62]
    Iron oxide nanoparticles enhancing magnetic resonance imaging
    This review explores the advancements in iron oxide nanoparticles (IONPs) as MRI contrast agents, emphasizing their synthesis, surface engineering, and impact ...
  63. [63]
    From Renewable Origin to Green Enhanced Oil Recovery Applications
    Scientific attempts to produce bio-based surfactants have produced promising results and some biosurfactants capable of replacing conventional chemical ...
  64. [64]
    Robust enzyme-iron oxide nanoparticle composites entrapped in ...
    Apr 1, 2024 · This is the first demonstration of the co-encapsulation of MNPs and enzymes in mesoporous silica nanospheres to form stabilized and immobilized enzyme systems.
  65. [65]
    Lipid nanoparticles for RNA delivery: Self-assembling vs driven ...
    A micro-emulsion approach was used to obtain CaP nanoparticles encapsulating siRNA, which were then mixed with cationic liposomes to form hybrid lipid-CaP [195] ...
  66. [66]
    “World in motion” – emulsion adjuvants rising to meet the pandemic ...
    Dec 21, 2021 · Emulsion adjuvants allow for antigen dose sparing, more rapid immune responses, and enhanced quality and quantity of adaptive immune responses.
  67. [67]
    Microemulsion phase behavior based on biodiesel and its ...
    Biodiesel-based microemulsions have unique advantages in the treatment of petroleum-contaminated soils, thanks to their ultra-low interfacial tension and strong ...
  68. [68]
    Advances and applications of supercritical carbon dioxide ...
    Combining scCO2 with microemulsion systems not only expands its solubilization range but also enhances the catalytic and separation efficiencies. When further ...Missing: capture | Show results with:capture
  69. [69]
    Ascendia Addresses Challenges Associated Emulsion and ...
    Aug 15, 2024 · Manufacturing them at a production scale sounds simple but often times can prove to be challenging. Using state-of-the-art manufacturing systems ...
  70. [70]
    Development of artificial neural network models to predict the ...
    May 16, 2024 · An Artificial Neural Network to predict the concentration range of microemulsion formation was developed using weight fraction, hydrophilic-lipophilic balance ...
  71. [71]
    Evaluation of Machine Learning Assisted Phase Behavior Modelling ...
    This paper evaluates the ability of machine learning (ML) algorithms to capture and reproduce complex multiphase behavior in surfactant–oil–water systems.
  72. [72]
    Transparent Water-in-Oil Dispersions: the Oleopathic Hydro-Micelle
    Transparent Water-in-Oil Dispersions: the Oleopathic Hydro-Micelle. T. P. HOAR &; J. H. SCHULMAN. Nature volume 152, pages 102–103 ( ...
  73. [73]
    Microemulsions Explained - Prospector Knowledge Center
    Sep 25, 2015 · Makeup removers represent one of the main applications for microemulsions. The term “microemulsion” was first used by Schulman1 in 1959 to ...
  74. [74]
    Winsor classification of emulsions (Winsor 1948) - ResearchGate
    Dispersions with the Winsor III phase are isotropic, thermodynamically stable, transparent to the eye, and have a very low interfacial tension(Friberg et al., ...
  75. [75]
    Equilibrium bicontinuous structure - Nature
    Sep 9, 1976 · SCRIVEN, L. Equilibrium bicontinuous structure. Nature 263, 123–125 (1976). https://doi.org/10.1038/263123a0. Download citation. Received: 02 ...
  76. [76]
    Nanoemulsions versus microemulsions: terminology, differences ...
    Dec 22, 2011 · This article reviews the differences and similarities between nanoemulsions and microemulsions in terms of their compositions, structure, ...
  77. [77]
    The HLD‐NAC Model for Mixtures of Ionic and Nonionic Surfactants
    Sep 28, 2008 · The HLD-NAC model has been used as an “equation of state” to predict the properties of microemulsion (μE) systems formulated with either anionic or nonionic ...
  78. [78]
    Nano-emulsions and micro-emulsions: clarifications of the ... - PubMed
    Nov 6, 2010 · Following these clarifications, we show how the confusion between nano-emulsions and micro-emulsions appears in the literature and how most of ...
  79. [79]
    Surfactant-free microemulsions - ScienceDirect.com
    ... micro-domain. We first discuss and compare the phase behaviour of these different types, starting from the classical microemulsion made from a flexible ...Missing: terminology | Show results with:terminology