Fact-checked by Grok 2 weeks ago

Micellar solubilization

Micellar solubilization is the process by which amphiphilic molecules, known as , self-assemble above their (CMC) to form colloidal aggregates called micelles in aqueous media, thereby incorporating poorly water-soluble (hydrophobic) substances into the micellar structure and enhancing their apparent without altering the solution's thermodynamic stability. This phenomenon relies on the reversible interaction between the solubilizate and the micelles, resulting in an isotropic solution where the solubilized material exhibits reduced thermodynamic activity compared to its pure form. The process involves partitioning of the solute between the bulk aqueous phase and the micellar pseudo-phase, driven by hydrophobic, electrostatic, and hydrogen bonding interactions. aggregate to form micelles with a hydrophobic core and hydrophilic shell, where the solubilizate location varies by polarity. Micellar solubilization has diverse applications, including in and detergents, pharmaceuticals (where it aids of poorly soluble drugs constituting approximately 70–90% of new candidates, improving and ), and .

Fundamentals

Definition and Overview

Micellar solubilization refers to the process of incorporating non-polar or poorly water-soluble substances, known as solubilizates, into micelles formed by surfactants in aqueous solutions, thereby enhancing the apparent solubility of these substances beyond their intrinsic solubility limits. This phenomenon relies on the self-assembly of amphiphilic surfactant molecules above a critical concentration, creating colloidal aggregates that can encapsulate hydrophobic molecules within their structure. The key components involved include the solvent, typically water; the association colloid, which is the surfactant capable of forming micelles; and the solubilizate, a hydrophobic or lipophilic molecule that partitions into the micellar assembly. Unlike simple , which results in a true molecular where the solute concentration is limited by , micellar solubilization produces a colloidal that allows total solubilizate concentrations to exceed the normal threshold through partitioning into the micelles. This distinction arises because the solubilizate is not fully molecularly dispersed but is instead associated with the micellar phase, forming a thermodynamically stable system without . The concept of micellar solubilization traces its origins to early investigations of soap solutions in the 1930s, where researchers observed the enhanced solubility of organic compounds within colloidal aggregates formed by soaps. A seminal study in 1937 demonstrated internal solubility in soap micelles, highlighting how these structures could dissolve substances otherwise insoluble in water. The term and underlying principles were further formalized in the mid-20th century, with influential reviews and texts in the 1950s solidifying its place in colloid and surface chemistry. In a basic representation, micelles are spherical aggregates with a hydrophobic core composed of tails, surrounded by a layer of partially exposed tails and heads, and an outer hydrophilic shell interfacing with water; solubilizates may localize in for highly hydrophobic compounds, the layer for moderately polar ones, or the shell for amphiphilic , depending on their molecular characteristics.

Surfactants and Micelle Formation

are amphiphilic molecules consisting of a hydrophilic head group, which is typically polar or ionic, and a hydrophobic tail composed of a non-polar alkyl chain. This dual nature allows to interact with both polar solvents like and non-polar substances, facilitating their surface-active properties. are classified into four main types based on the charge of their head group: anionic, cationic, non-ionic, and zwitterionic. Anionic , such as (, CH₃(CH₂)₁₁OSO₃⁻Na⁺), possess a negatively charged head group and are widely used due to their strong cleaning ability. Cationic , like cetyltrimethylammonium bromide (CTAB, CH₃(CH₂)₁₅N(CH₃)₃⁺Br⁻), feature a positively charged head and often exhibit properties. Non-ionic , exemplified by Tween 80 (a polyoxyethylene with the general structure involving chains attached to a ester), lack electrical charge and are valued for their mildness and compatibility with other . Zwitterionic , such as 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (, featuring both a and a anion in the head group), contain both positive and negative charges and are known for their pH-independent behavior and . The (CMC) represents the threshold concentration above which micelles begin to form in , marking a sharp change in physical properties like or conductivity. For homologous series of , the CMC can be approximated using the Klevens : \log(\text{CMC}) = A - B \times n where n is the number of carbon atoms in the hydrophobic chain, and A and B are empirical constants dependent on the type (e.g., B \approx 0.3 for many ionic ). This relationship highlights how longer hydrophobic chains lower the CMC by enhancing the . Micelle formation occurs through the self-assembly of molecules in , driven primarily by the , which minimizes the system's by sequestering non-polar tails away from water. Below the , surfactants exist as monomers or adsorb at interfaces; above the , they aggregate into to reduce unfavorable interactions between hydrophobic tails and water molecules. The resulting morphology—spherical, cylindrical, or vesicular—depends on factors such as the surfactant's packing parameter (\rho = v / (a \cdot l), where v is the tail volume, a the head area, and l the tail length), surfactant concentration, and environmental conditions; for instance, spherical predominate when \rho \leq 1/3, cylindrical when $1/3 < \rho \leq 1/2, and vesicular or bilayer structures when $1/2 < \rho \leq 1. Thermodynamically, micelle formation is predominantly entropy-driven, arising from the release of structured water molecules that previously surrounded the hydrophobic tails, thereby increasing the overall entropy of the system despite a sometimes endothermic enthalpy change. This entropy gain from disrupted water cages outweighs any enthalpic penalties, making the process spontaneous above the under typical conditions.

Mechanism

Micelle Structure and Solubilizate Partitioning

Micelles, formed by the self-assembly of amphiphilic surfactants above the (CMC), exhibit a structured architecture that facilitates solubilization. The micelle consists of three primary regions: a hydrophobic core composed of the aggregated hydrocarbon tails of the surfactant molecules, which excludes water; a palisade layer at the interface between the tails and heads, characterized by partial hydration and intermediate polarity; and an outer hydrophilic shell formed by the polar head groups in direct contact with the aqueous environment. This radial organization creates distinct microenvironments that dictate the partitioning of solubilizates based on their molecular polarity. Solubilizates partition into specific sites within the micelle depending on their hydrophilicity or hydrophobicity. Non-polar solubilizates, such as hydrocarbons like or oils, preferentially locate in the hydrophobic core due to favorable van der Waals interactions with the surfactant tails. Polar or amphiphilic solubilizates, exemplified by short-chain alcohols, tend to reside in the palisade layer or at the micelle surface, where they interact with the polar head-tail interface or the hydrated shell. This site-specific incorporation enhances the overall solubility of poorly water-soluble compounds by shielding non-polar moieties from the aqueous bulk. The extent of partitioning is quantified by the partition coefficient K, defined as the ratio of the solubilizate concentration in the micelle to that in the bulk aqueous phase: K = \frac{[S]_m}{[S]_w}, where [S]_m and [S]_w represent the concentrations in the micellar and aqueous phases, respectively. Values of K increase with the hydrophobicity of the solubilizate, reflecting stronger affinity for the micellar interior over the aqueous environment. Solubilization capacity refers to the maximum amount of solubilizate that can be accommodated per micelle, often expressed as the molar solubilization capacity \chi, the moles of solubilizate per mole of micellar surfactant. Above the CMC, solubilization capacity increases linearly with surfactant concentration, as additional micelles provide more incorporation sites without altering the aggregation number significantly. For instance, non-ionic surfactants like polysorbates can solubilize up to several moles of hydrophobic drugs per mole of surfactant, depending on chain length. Experimental determination of partitioning and capacity typically employs techniques such as ultraviolet (UV) spectroscopy, which measures changes in absorbance due to the micellar environment, or equilibrium dialysis, where solubilizate distribution across a semi-permeable membrane separating micellar and bulk phases is analyzed. These methods allow precise quantification of K and \chi by varying surfactant concentrations and monitoring solubility enhancements.

Kinetic and Thermodynamic Processes

Micellar solubilization is governed by thermodynamic principles that favor the incorporation of hydrophobic solubilizates into micelles, primarily through a negative change in Gibbs free energy (). This process is described by the equation ΔG = ΔH - TΔS, where ΔH is the enthalpy change, T is the temperature, and ΔS is the entropy change. The driving force arises from the hydrophobic effect, which minimizes unfavorable interactions between nonpolar solubilizate molecules and water, coupled with a reduction in interfacial tension at the micelle-water boundary, rendering ΔG negative and spontaneous under typical conditions. Equilibrium in micellar systems can be modeled using either the phase separation approach, which treats micelles as a distinct phase analogous to macroscopic separation, or the mass action model, which views micelle formation as a chemical equilibrium between surfactant monomers and aggregates. The mass action model is more consistent with experimental observations of surfactant solubility and solubilizate uptake above the , as it accounts for the gradual increase in aggregate concentration without invoking phase discontinuities. In this framework, the micelle association constant is given by K_m = \frac{[M]}{[S]^n} where [M] is the concentration of micelles, [S] is the concentration of surfactant monomers, and n is the aggregation number representing the average number of monomers per micelle. The kinetics of solubilization involve distinct mechanisms depending on the solubilizate's interaction with the micelle. Bulk solubilization occurs via diffusion of the solubilizate into the hydrophobic core of the micelle, while surface solubilization proceeds through adsorption at the polar headgroup interface or palisade layer. These processes often follow pseudo-first-order kinetics, with the rate of incorporation proportional to the surfactant concentration above the critical micelle concentration, as described by rate equations where the observed rate constant k_obs reflects the mean lifetime of the solubilizate in solution before micelle entry. The aggregation number, typically ranging from 20 to 100 surfactant molecules per micelle in spherical aggregates, plays a key role in solubilization stability by determining the micelle's capacity to accommodate solubilizates without structural disruption. This parameter is commonly determined using light scattering techniques, which measure changes in scattered intensity to infer micelle size and polydispersity, providing values that correlate with solubilization efficiency in systems like alkylpyridinium surfactants. Energy barriers for solubilizate entry into micelles are generally low, on the order of 30 kJ/mol, due to the dynamic fluctuations of micelle structure that facilitate transient openings in the surfactant shell. These fluctuations enable rapid solute exchange, with activation energies derived from temperature-dependent kinetic studies aligning with diffusion-limited processes rather than high-barrier rearrangements.

Applications

Cleaning and Detergents

Micellar solubilization is fundamental to the efficacy of detergents in cleaning applications, where surfactants reduce the interfacial tension between water and hydrophobic soils, enabling the formation of micelles that encapsulate grease, oils, and dirt particles within their hydrophobic cores. This encapsulation lifts soils from surfaces and suspends them in the wash solution, preventing redeposition and ensuring thorough cleaning. The process relies on surfactant aggregation above the (CMC), as briefly referenced in the mechanisms of micelle formation. In laundry applications, anionic such as linear alkylbenzene sulfonates (LAS) dominate formulations due to their strong soil removal capabilities, particularly for particulate and oily stains on fabrics. These form micelles that effectively solubilize and emulsify soils, with micelle size playing a key role in efficiency—larger micelles can accommodate more hydrophobic material, enhancing removal rates for embedded dirt. Typical laundry detergent formulations incorporate 10-20% total to achieve concentrations well above the CMC during use, optimizing cleaning performance while balancing cost and stability. For instance, combinations of anionic and nonionic improve overall detergency by targeting both particulate soils and greasy residues. For hard surface cleaning, such as in kitchen or bathroom products, micellar solubilization facilitates the emulsification of oils and fats, allowing them to be dispersed in aqueous solutions without leaving residues. This mechanism outperforms traditional soaps, which precipitate as insoluble calcium or magnesium salts in hard water, forming scum and reducing cleaning efficiency; synthetic detergents, by contrast, maintain micelle integrity and solubilization in such conditions. Household cleaners often leverage spontaneous emulsification driven by micelles to rapidly remove oily soils from tiles or countertops. Efficiency in these applications is often evaluated using the solubilization index, which quantifies the amount of oil or soil removed per unit mass of surfactant, highlighting how micelle capacity directly correlates with cleaning power. This metric underscores the advantages of synthetic detergents developed since the 1930s, when pioneers like Shell's Teepol introduced alkyl sulfates that exploited micellar solubilization for superior performance over soaps, especially in wartime shortages of natural fats. Branched alkylbenzene sulfonates, commercialized in the early 1930s, further advanced this by enabling effective oil removal without precipitation issues. Environmental considerations have driven the shift toward biodegradable surfactants in modern detergents to mitigate aquatic impacts, as non-degradable types can persist in wastewater and harm ecosystems. Readily biodegradable anionics, such as those meeting OECD standards, break down rapidly under aerobic conditions, reducing toxicity to aquatic organisms and soil accumulation from laundry effluents. This focus ensures that micellar-based cleaning maintains efficacy while minimizing long-term ecological footprints.

Pharmaceuticals and Drug Delivery

Micellar solubilization plays a crucial role in pharmaceuticals by enhancing the aqueous solubility of poorly water-soluble drugs, particularly those classified as Biopharmaceutics Classification System (BCS) Class II or IV, which exhibit low solubility but high permeability. For instance, glimepiride can achieve solubility increases of up to 430-fold when incorporated into micelles formed by surfactants such as sodium dodecyl sulfate (SDS) or cetyltrimethylammonium bromide (CTAB) mixed with non-ionic agents like Tween-80, while paclitaxel formulations using polymeric micelles can achieve increases of thousands-fold. This enhancement occurs as hydrophobic drug molecules partition into the micelle core, enabling higher drug loading and preventing precipitation in aqueous media. Common formulation types include micellar solutions, gels, and nano-sized polymeric micelles, which offer stability and controlled release. Notable FDA-approved examples are Genexol-PM, a poly(ethylene glycol)-poly(D,L-lactide) micellar formulation of approved for breast and lung cancer treatment, and Restasis (cyclosporine A in castor oil nanoemulsion) for ocular dry eye syndrome. These systems facilitate various delivery routes: intravenous (IV) for systemic chemotherapy, oral for improved gastrointestinal absorption, and topical for localized therapy. Micelles protect encapsulated drugs from enzymatic degradation and enhance permeability across biological barriers, such as the intestinal epithelium or skin, by mimicking lipid bilayers and promoting endocytosis. Despite these benefits, challenges persist, including the potential toxicity of ionic surfactants like SDS, which can cause hemolysis or irritation at high concentrations, limiting their use in parenteral formulations. Advances since the early 2000s have shifted toward non-ionic polymeric micelles, such as those based on (poloxamer) block copolymers, which exhibit lower cytotoxicity and enable targeted delivery by inhibiting P-glycoprotein efflux pumps in multidrug-resistant cells. Pharmacokinetically, micellar partitioning influences drug release rates, with slower dissociation from the core leading to sustained profiles; in vivo studies demonstrate increased area under the curve (AUC) values, such as 4.9- to 5.7-fold for in polymeric micelles compared to free drug, thereby improving bioavailability and therapeutic efficacy. Recent developments include , a micellar paclitaxel formulation in phase III trials as of 2023 for advanced solid tumors, demonstrating improved efficacy in multidrug-resistant cancers.

Environmental Remediation

Micellar solubilization plays a crucial role in environmental remediation by leveraging surfactants to disperse and mobilize hydrophobic pollutants in uncontrolled natural systems. In oil spill cleanup, dispersants such as form micelles that reduce the interfacial tension between oil and water, breaking large oil slicks into small droplets typically less than 100 μm in diameter. This process enhances the oil's surface area, facilitating greater exposure to water and promoting biodegradation by marine microorganisms. During the , approximately 1.84 million gallons of Corexit were applied, both on the surface and subsea, which increased the oil-water interface by up to 1000 times and aided in dispersing the oil into the water column for microbial degradation. For soil and groundwater remediation, surfactant flushing mobilizes persistent organic pollutants like polycyclic aromatic hydrocarbons (PAHs) and chlorinated solvents through micelle formation, which solubilizes these hydrophobic compounds and increases their aqueous solubility. Nonionic surfactants such as have demonstrated up to 62% removal of PAHs like phenanthrene in contaminated soils over 15 days, while anionic surfactants like achieve 97% removal of diesel components in field applications. In pump-and-treat systems, surfactants are injected via wells to flush contaminants from the subsurface, followed by extraction and above-ground treatment, proving effective for dense non-aqueous phase liquids (DNAPLs) such as , with removal rates reaching 95% in pilot-scale operations at sites like Alameda Point. The Deepwater Horizon incident exemplifies the scale of micellar dispersant application, where subsea injection of Corexit created droplets of 10-100 μm, preventing resurfacing and enhancing dilution across vast water volumes, though only about 3% of the total oil was mechanically recovered by skimmers. Ecological impacts of micellar dispersants are dual-edged: they accelerate oil degradation by increasing bioavailability to microbes, thereby reducing long-term surface contamination and protecting coastal habitats, as seen in faster recovery of some intertidal zones post-spill. However, dispersed oil droplets can exhibit heightened toxicity to subsurface marine life, including fish embryos, larvae, and corals, with synergistic effects amplifying harm when mixed with oil. These concerns prompted regulations following the 1989 Exxon Valdez spill, including the Oil Pollution Act of 1990, which improved spill response planning and dispersant pre-approval processes, and post-Deepwater Horizon updates to the National Contingency Plan in 2023, mandating toxicity testing and limiting applications to protect sensitive ecosystems. Emerging applications include micellar-enhanced ultrafiltration (MEUF) for wastewater treatment, where surfactants form micelles to solubilize organic contaminants like pharmaceuticals and dyes, achieving retention rates over 90% via ultrafiltration membranes with higher flux than traditional methods. This technique targets emerging pollutants in industrial effluents, such as naphthenic acids from oil sands, and allows surfactant recovery for reuse, making it a sustainable option for large-scale water purification.

Influencing Factors

Surfactant Properties and Concentration

The properties of surfactants, particularly the length of their hydrophobic tails and the nature of their hydrophilic head groups, play a pivotal role in determining the efficiency of micellar solubilization. Longer hydrophobic chain lengths generally lower the (CMC) by enhancing the hydrophobic effect that drives micelle formation, as each additional methylene group in the tail reduces the CMC by approximately a factor of 2. Furthermore, extended tails increase the volume of the micellar hydrophobic core, thereby enhancing the solubilization capacity for non-polar solubilizates, with studies showing a linear increase in solubility as tail length rises from C8 to C16 for nonionic surfactants like alkyl polyglucosides. In contrast, the head group influences micelle stability through steric and electrostatic interactions; ionic head groups, such as sulfate in (SDS), elevate the CMC compared to nonionic counterparts due to electrostatic repulsion between charged heads, which hinders aggregation unless screened by counterions. Surfactant concentration directly governs solubilization capacity, with no significant micelle formation or solubilization occurring below the , where monomers predominate in solution. Above the , the amount of solubilized material increases linearly with surfactant concentration, proportional to (C - ), where C is the total surfactant concentration, reflecting the growing number of micelles available for partitioning. This relationship holds for both ionic and nonionic surfactants, though the slope (solubilization capacity) varies with molecular structure, typically ranging from 0.1 to 1 mole of solute per mole of surfactant for hydrophobic drugs like ibuprofen. Blending surfactants to form mixed micelles often yields synergistic effects that enhance solubilization efficiency, particularly when combining anionic and nonionic types, such as with Tween 80, where favorable packing and reduced electrostatic repulsion lower the overall compared to individual components. This synergy arises from favorable intermolecular interactions that stabilize the mixed aggregates, lowering the free energy of micellization and increasing the effective core volume for solubilizates without altering the fundamental amphiphilic nature of the surfactants. The aggregation number, which denotes the average number of surfactant molecules per micelle, typically increases with rising surfactant concentration above the CMC, leading to larger micelles that can impact stability by altering polydispersity and potential for secondary aggregation. Small-angle neutron scattering (SANS) studies on cationic surfactants like cetyltrimethylammonium bromide (CTAB) confirm this trend, showing aggregation numbers rising from around 50 at near-CMC concentrations to over 100 at 10 times the CMC, which enhances solubilization for non-polar probes but may reduce kinetic stability in dilute systems. Selection of surfactants for micellar solubilization often relies on the hydrophilic-lipophilic balance (HLB) value, with optimal ranges of 8-18 favoring oil-in-water systems where solubilization predominates, as higher HLB values promote micelle formation in aqueous media while maintaining sufficient hydrophobicity for core accommodation of lipophilic guests. For instance, nonionic surfactants like polysorbate 80 (HLB ≈ 15) exemplify this range, exhibiting superior solubilization of poorly water-soluble compounds compared to those with HLB below 8, which favor water-in-oil partitioning.

Environmental Conditions

Micellar solubilization is significantly influenced by temperature, which affects both the formation of micelles and the partitioning of solubilizates within them. For non-polar compounds, higher temperatures generally enhance solubilization by promoting micellar growth and expanding the hydrophobic core, thereby increasing the accommodation capacity for hydrophobic molecules. This effect is particularly pronounced in systems with block copolymer surfactants like , where temperature-induced dehydration of hydrophilic chains drives core expansion and solubilization of otherwise insoluble drugs. However, for non-ionic surfactants, temperatures exceeding the cloud point lead to phase separation and micelle disruption, reducing overall solubilization efficiency. The pH of the solution plays a critical role in modulating micellar solubilization, especially for ionic surfactants where head group ionization is pH-dependent. In carboxylate-based anionic surfactants, low pH causes protonation of the carboxylate groups, increasing electrostatic repulsion and raising the critical micelle concentration (CMC), which in turn diminishes micelle formation and solubilization capacity. Most micellar systems exhibit optimal performance at neutral to slightly alkaline pH levels of 7-9, where surfactant ionization is balanced to minimize repulsion while maintaining stability. For instance, in amphoteric surfactants, pH adjustments can reversibly control solubilization, with basic conditions enhancing drug loading by altering head group charge. Ionic strength, modulated by added salts, impacts micellar solubilization by screening electrostatic repulsions between charged head groups, thereby lowering the and facilitating micelle assembly at lower surfactant concentrations. This effect follows the , where kosmotropic ions (e.g., sulfate) more effectively promote micellization and enhance solubilization compared to chaotropic ions (e.g., thiocyanate), due to their stronger structuring of the surrounding water and better stabilization of the micellar interface. In high-salt environments, this screening can increase solubilizate partitioning into the micelle core, particularly for ionic surfactants. Responses to these environmental conditions vary by solubilizate type, with polar compounds showing greater sensitivity to pH changes due to alterations in their ionization state and interactions with the micellar surface. For example, phenolic solubilizates exhibit pH-dependent solubility isotherms, where acidic conditions reduce partitioning into micelles of by promoting unionized forms that favor the aqueous phase, while neutral pH enhances incorporation via hydrogen bonding or electrostatic interactions. Non-polar solubilizates, in contrast, are less affected by pH but respond more to temperature and ionic strength through changes in micelle hydrophobicity. In practical applications, precise control of temperature is essential in pharmaceutical formulations to optimize drug loading without exceeding stability limits, as elevated temperatures can improve solubility but risk micelle destabilization in storage. Similarly, pH adjustments are crucial in environmental remediation, such as using alkaline conditions (pH >8) with anionic surfactants to enhance oil spill cleanup by improving emulsification and solubilization of hydrocarbons while countering acidity. These optimizations ensure efficient solubilization under varying field or processing conditions.

References

  1. [1]
  2. [2]
    Micellar Solubilization - an overview | ScienceDirect Topics
    Micellar solubilization also called as solubilization by surfactants (Fig. 2.7), are the molecules with prominent polar and non-polar portions.
  3. [3]
    Micelles transforming modern medicine: a comprehensive review on ...
    Sep 22, 2025 · Micelles enhances the solubility and bioavailability of biologically active compounds by providing an ideal structure for entrapment and ...Missing: definition | Show results with:definition
  4. [4]
    Micellar Solubilization of Some Poorly Soluble Antidiabetic Drugs
    Micellar solubilization is a widely used alternative for the dissolution of poorly soluble drugs (12–14). Depending upon the drug hydrophobicity, it can be ...
  5. [5]
    Micellar solubilization of drugs. - University of Alberta
    Micellar solubilization is a powerful alternative for dissolving hydrophobic drugs in aqueous environments. In this work, we provide an insight into this ...
  6. [6]
    Internal solubility in soap micelles - RSC Publishing
    Internal solubility in soap micelles. A. S. C. Lawrence, Trans. Faraday Soc., 1937, 33, 815 DOI: 10.1039/TF9373300815 ... Education in Chemistry · Open Access ...
  7. [7]
    Micellar solubilization – Knowledge and References - Taylor & Francis
    Micellar solubilization ... Two early but very useful references on solubilization are McBain and Hutchinson's book (1955) and a review article by Klevens (1950).
  8. [8]
    Self-assembly of surfactants: An overview on general aspects of ...
    Surfactants are normally amphiphilic molecules with a long hydrocarbon tail and a polar head group. The head group may be anionic, cationic or nonionic.
  9. [9]
    Surfactants: physicochemical interactions with biological ... - PMC
    Feb 3, 2021 · Surfactants are amphiphilic molecules capable of reducing the surface tension between two immiscible phases (Otzen 2017). These molecules ...
  10. [10]
    Zwitterionic Surfactant - an overview | ScienceDirect Topics
    Table 1 shows the structures of the more common zwitterionic surfactants. The positively charged group is usually a quaternary ammonium or an imidazolium group.
  11. [11]
    Critical Micelle Concentration - an overview | ScienceDirect Topics
    Critical micelle concentration (CMC) is defined as the surfactant concentration at which micelle formation is first seen in the solution. The values of CMC ...
  12. [12]
    Soft Matter - RSC Publishing
    When at least one of the amino groups of. CmN2N was protonated the logarithm of the critical micelle concentration (cmc) followed the well- known Klevens ...
  13. [13]
    (PDF) The Effect of Hydrocarbon Chain Length on the Critical ...
    Aug 7, 2025 · It was observed that CMC decreases with increasing hydrocarbon chain length as hydrophobicity increases with chain length thus resulting in the ...
  14. [14]
    Micelle Formation and the Hydrophobic Effect - ResearchGate
    Aug 7, 2025 · The tendency of amphiphilic molecules to form micelles in aqueous solution is a consequence of the hydrophobic effect.
  15. [15]
    Surfactant Self-Assembling and Critical Micelle Concentration
    May 6, 2020 · Critical micelle concentration (CMC) is the main chemical–physical parameter to be determined for pure surfactants for their ...
  16. [16]
    Cylindrical Micelle - an overview | ScienceDirect Topics
    As a general rule, spherical micelles are formed when ρ ≤ ⅓, cylindrical micelles are formed at ⅓ < ρ ≤ ½ and membranes arise when ½ < ρ ≤ 1. As shown in Fig. 1 ...
  17. [17]
    Molecular dynamics simulation of four typical surfactants in aqueous ...
    Jan 24, 2019 · The mainly force for micellization process is entropy-driven, and as the temperature increases, the entropy-driven contribution is gradually ...
  18. [18]
    Micelle Formation and the Hydrophobic Effect - ACS Publications
    This paper concerns the formation of micelles, which are the simplest form of amphiphilic assemblies. Our treatment of this phenomenon is ...Missing: driven | Show results with:driven
  19. [19]
    [PDF] Micellar solubilization of drugs. - University of Alberta
    Jul 8, 2005 · In this work, we provide a review of micellar solubilization of drugs in surfactant systems, blending it with basic information on ...
  20. [20]
    Solubilization. | Chemical Reviews - ACS Publications
    Micellar solubilization in aqueous binary surfactant systems: Barbituric acids in mixed anionic + nonionic or cationic + nonionic mixtures. Journal of ...
  21. [21]
    A unified phenomenological theory for micellization and solubilization
    A common basis has been chosen such that the Gibbs free energies of formation can be written in all cases as the sum of hydrophobic effect and the stretching- ...
  22. [22]
    Examination of micelle formation by phase rule - ScienceDirect
    Two typical models of micelle formation, the mass action model and the phase separation model, have been examined to see which model is consistent with the ...
  23. [23]
    A comparison of micellar aggregation numbers as determined by ...
    A comparison of micellar aggregation numbers as determined by dye solubilization and light scattering. Author links open overlay panel. Paul T Jacobs
  24. [24]
    None
    Below is a merged summary of all provided segments on "Kinetics of Solubilization, Energy Barriers, and Activation Energy for Solute Entry into Micelles." To retain all information in a dense and organized manner, I will use a combination of narrative text and tables in CSV format where appropriate. The response consolidates details from all segments, avoiding redundancy while ensuring completeness.
  25. [25]
    Solubilization - an overview | ScienceDirect Topics
    The process of solubilization is also important in detergency, whereby fats and oils are removed by incorporation into the hydrocarbon core of the micelles.
  26. [26]
    Laundry Detergents: An Overview - J-Stage
    5 MECHANISM OF SOIL REMOVAL BY LAUNDRY. DETERGENT/SURFACTANT. Surfactants can work in three different ways: roll-up, emulsification, and solubilization. 5.1 ...
  27. [27]
    Design and performance optimisation of detergent product ...
    In a detergent containing both nonionic and anionic surfactant, the anionic surfactant contributes to cleaning performance in soil removal and nonionic ...
  28. [28]
    Insights on Acrylic Polymer Films Dewetting and Artificial Soil Removal
    May 12, 2020 · Micelles' size is almost unaltered after the surfactant interaction with soil, and the volume fraction decreases for both surfactants, similarly ...
  29. [29]
    US6376446B1 - Liquid detergent composition - Google Patents
    The surfactant (nonionic and anionic) concentrations in most commercially available concentrates range from about 15% to about 26%. Typically, blends of ...<|separator|>
  30. [30]
    Spontaneous Emulsification of Triolein Induced by Mixed Micellar ...
    Spontaneous emulsification can reduce the time required to remove oily soils from hard surfaces and enhance the detergency, because this type of ...
  31. [31]
    Pathways of Membrane Solubilization: A Structural Study of Model ...
    Mar 9, 2023 · This study used small-angle X-ray scattering to determine the structures of lipid/detergent aggregates at different ratios and temperatures.
  32. [32]
    The biodegradation of surfactants in the environment - ScienceDirect
    Cationic surfactants, though readily biodegradable in aerobic environments, are toxic even at low concentrations. Therefore application to agricultural soil ...
  33. [33]
    Environmental risks and toxicity of surfactants - NIH
    Sep 29, 2021 · This article presents a compendium of knowledge on environmental toxicity and risks, analytical techniques, and remediation methods of surfactants
  34. [34]
    In vivo evaluation of polymeric micellar paclitaxel formulation
    Aug 7, 2025 · The aqueous solubility of the formulation prepared by Dabholkar et al. increased 5000 times, with a concentration of 5 mg/mL, while the other ...<|control11|><|separator|>
  35. [35]
    Micellar solubilization of drugs - PubMed
    Jul 8, 2005 · Purpose: Micellar solubilization is a powerful alternative for dissolving hydrophobic drugs in aqueous environments.
  36. [36]
    Polymeric Micelles in Anticancer Therapy: Targeting, Imaging ... - NIH
    Currently, five micellar formulations for anticancer therapy are under clinical evaluation, of which Genexol-PM has been FDA approved for use in patients with ...
  37. [37]
    Polymeric micelles in drug delivery: An insight of the techniques for ...
    Apr 10, 2021 · Polymeric micelles represent a promising strategy to overcome several issues related to drug delivery.
  38. [38]
    Exploring the Application of Micellar Drug Delivery Systems in ... - NIH
    The Food and Drug Administration (FDA) has authorized various formulations incorporating PLA, further demonstrating its suitability for rapid clinical ...
  39. [39]
    Production and Cytotoxicity of Solid Lipid Nanoparticles (SLNs) - MDPI
    Indeed, it has been documented that non-ionic surfactants are less toxic, than ionic surfactants [3,15], while cationic surfactants have been reported to ...<|control11|><|separator|>
  40. [40]
    Drug Delivery Systems Based on Pluronic Micelles with ... - NIH
    Jul 25, 2022 · These micellar systems can be an efficient alternative for the treatment of bacterial oral diseases, being suitable either by injection or by a topical ...
  41. [41]
    Polymeric micelles for the solubilization and delivery of cyclosporine A
    Similar to the pharmacokinetic study in these rats, the polymeric micellar formulation gave rise to 5.7 and 4.9-fold increase in the AUC of CsA in blood and ...
  42. [42]
    Polymeric Micelles for the Delivery of Poorly Soluble Drugs - PMC
    That is, micelles with a higher PEG density and a brush-like conformation had increased AUC in vivo which is essential for improved efficacy of polymeric ...
  43. [43]
    Dispersants as an oil spill clean-up technique in the marine ... - NIH
    This review discusses the application of dispersants onto oil spills as enhancement of microbial degradation within the water column.
  44. [44]
    The Use of Dispersants in Marine Oil Spill Response - NCBI Bookshelf
    The small dispersed oil droplets (generally 10-100 microns [µm]) remain suspended in the water column and are an indicator of dispersant effectiveness (Li et ...
  45. [45]
    What Have We Learned About Using Dispersants During the Next ...
    Apr 20, 2015 · During the Deepwater Horizon oil spill response, skimmers only managed to recover approximately 3 percent of the oil released. Dispersants ...
  46. [46]
    Soil Contaminants and Their Removal through Surfactant-Enhanced ...
    This review provides a comprehensive analysis of the effectiveness of surfactants in enhancing the remediation of contaminated soils.
  47. [47]
    Long-term ecological impacts from oil spills - PubMed Central - NIH
    Large scale persistent ecological effects included impacts to deep ocean corals, failed recruitment of oysters over multiple years, damage to coastal wetlands, ...
  48. [48]
    10. What Are Some Environmental Impacts from Dispersant Use?
    Both dispersants and dispersed oil particles are toxic to some marine organisms. But because concentrations of dispersed oil are quickly diluted in the ocean,
  49. [49]
    EPA Issues Final Rule Governing Use of Chemical Dispersants on ...
    Jun 13, 2023 · The new NCP continues to sanction broad scale use of chemical dispersants, but now it also requires that the decision to use dispersants in any ...
  50. [50]
    Water and wastewater treatment by micellar enhanced ultrafiltration
    Micellar-enhanced ultrafiltration (MEUF) is a surfactant-based method used to remove low concentrations of contaminants from water and wastewater streams.
  51. [51]
    Effect of alkyl chain length on solubility and micelle formation
    Aqueous solubility and critical micelle concentration (CMC) of alkylsulfonates were measured for surfactants with different alkyl chain lengths, and with ...Missing: solubilization | Show results with:solubilization
  52. [52]
    Molecular behavior of ionic and nonionic surfactants in saline medium
    From the obtained results, it can be concluded that the ionic surfactants, due to electrostatic repulsion between ionic head groups, showed higher cmc values ...
  53. [53]
    Micellar solubilization of poorly water-soluble drugs
    Abstract. Objective: This study aims to clarify the role of surfactant and drug molecular structures on drug solubility in micellar surfactant solutions.Missing: key components
  54. [54]
    Investigating Synergism in Critical Micelle Concentration of Anionic ...
    The Critical Micelle Concentrations (CMCs) of mixed surfactants were sharply lower than that of sole anionic surfactant (SDS or SDBS). With the increase in mole ...Introduction · Materials And Methods · Results And Discussion
  55. [55]
    Synergistic effects between a non-ionic and an anionic surfactant on ...
    Dec 20, 2023 · The synergism, at mixed monolayers or in mixed micelles, which arises from interactions between differently charged surfactant headgroups (non-ionic/ionic or ...
  56. [56]
    Correlation between the geometrical shape and growth behaviour of ...
    Small tablet-shaped micelles formed by CTAB are found to grow only weakly in size with increasing surfactant concentration.
  57. [57]
    Small angle neutron scattering measurements of aggregation ...
    For pure cationic surfactants DTAB to HTAB, there is an increase in aggregation number and decrease in charge fraction, which indicates that, the micelle ...
  58. [58]
    Hydrophilic-Lipophilic Balance - an overview | ScienceDirect Topics
    Therefore, surfactants having the HLB values between 9 and 12 stabilize O/W emulsion, whereas those HLB values around 3–8 stabilize W/O emulsion [6].
  59. [59]
    HLB Scale, Solubilization, Detergency and Adsorption at Solid ...
    Based on the HLB value, the surfactant is classified as water-soluble or oil-soluble according to its relative solubility. A low number of hydrophilic groups ...