Fact-checked by Grok 2 weeks ago

Upper critical solution temperature

The upper critical solution temperature (UCST) is the critical temperature above which the components of a or multicomponent become completely miscible across all compositions, forming a single homogeneous phase, while below this temperature, into two immiscible phases occurs for certain composition ranges. This phenomenon arises in systems of partially miscible liquids or solutions where intermolecular attractions (such as hydrogen bonding or van der Waals forces) favor segregation at lower temperatures, but dominates above the UCST to promote mixing. The UCST marks the apex of the coexistence curve in a temperature-composition , typically near equimolar ratios (around 50:50 composition), and its value depends on factors like , molecular weight (especially in ), and the presence of additives or salts. In contrast to the (LCST), where occurs upon heating due to entropy-driven effects, the UCST behavior is enthalpically driven, with mixing favored at higher temperatures as endothermic interactions overcome exothermic attractions between like molecules. UCST systems are common in non-polar mixtures and solutions, and the transition can be tuned by varying chain length, concentration, or environmental conditions like or . For instance, the classic -cyclohexane system exhibits a UCST around 34°C (307 K) for low-molecular-weight , above which the fully dissolves in the . Another example is the hexane-nitrobenzene , which achieves complete upon heating to approximately 19°C (292 K). UCST transitions are particularly significant in for designing thermoresponsive materials, such as hydrogels for or smart coatings that switch with changes under physiological conditions. These properties enable applications in separation processes, like extracting components from mixtures by controlled cooling below the UCST, and in studying near the phase boundary, where fluctuations lead to enhanced transport properties. While many UCST systems operate at ambient or elevated s, recent advances have developed aqueous systems with tunable UCSTs near body temperature, expanding biomedical uses.

Fundamentals

Definition of UCST

The upper critical solution temperature (UCST) is the highest temperature at which a binary mixture of partially miscible substances exhibits into two coexisting liquid phases; above this temperature, the components become fully miscible across all compositions. This critical point marks the apex of the two-phase region in the temperature-composition , beyond which overcomes intermolecular forces that otherwise promote immiscibility. The UCST phenomenon was first observed in the late , notably in systems like phenol-aniline-water, where partial was noted under varying temperatures. Formal thermodynamic descriptions of such behavior emerged in the early , building on foundational work in solution theory to explain the conditions for criticality in mixtures. below the UCST is primarily enthalpy-driven, arising from unfavorable interactions between unlike molecules that outweigh the entropic gain from mixing, leading to a coexistence curve bounding the immiscible region. In contrast, UCST behavior is the inverse of (LCST) systems, where demixing occurs upon heating due to entropy-dominated effects. The UCST and corresponding critical composition are thermodynamically defined as the point where the second and third partial derivatives of the molar Gibbs free energy of mixing (g_{\text{mix}}) with respect to mole fraction vanish: \left( \frac{\partial^2 g_{\text{mix}}}{\partial x^2} \right)_{T,P} = 0, \quad \left( \frac{\partial^3 g_{\text{mix}}}{\partial x^3} \right)_{T,P} = 0 The Gibbs free energy of mixing typically takes the form G_{\text{mix}} = nRT (x_1 \ln x_1 + x_2 \ln x_2) + \text{interaction terms}, where the logarithmic terms represent the ideal entropic contribution, and interaction terms account for enthalpic effects.

Thermodynamic Principles

The upper critical solution temperature (UCST) arises from the thermodynamics of mixing, where the of mixing, \Delta G_{\text{mix}}, dictates phase stability in binary systems. This is expressed as \Delta G_{\text{mix}} = \Delta H_{\text{mix}} - T \Delta S_{\text{mix}}, with \Delta H_{\text{mix}} representing the enthalpic contribution and T \Delta S_{\text{mix}} the temperature-weighted entropic term. Below the UCST, \Delta G_{\text{mix}} > 0 due to unfavorable mixing energetics, promoting into two immiscible phases, whereas above the UCST, \Delta G_{\text{mix}} < 0, stabilizing the homogeneous mixed state. In UCST systems, the entropy of mixing \Delta S_{\text{mix}} approximates the ideal configurational value, which is positive and favors homogeneity by increasing disorder upon blending components. However, the enthalpy of mixing \Delta H_{\text{mix}} is typically positive, reflecting net repulsive interactions between dissimilar molecules that outweigh attractive forces at lower temperatures. This enthalpic dominance suppresses mixing until thermal energy amplifies the entropic contribution sufficiently to render \Delta G_{\text{mix}} negative. Phase separation boundaries in UCST systems are mapped by the binodal and spinodal curves on temperature-composition diagrams. The binodal curve traces the equilibrium coexistence of two phases, defined by the common tangent rule on the \Delta G_{\text{mix}} versus composition plot, ensuring equal chemical potentials across phases. Inside the binodal but outside the spinodal lies the metastable region, where nucleation is required for separation. The spinodal curve marks the boundary of local stability, occurring where the second derivative of the free energy with respect to composition is zero: \frac{\partial^2 \Delta G_{\text{mix}}}{\partial \phi^2} = 0, beyond which infinitesimal fluctuations amplify spontaneously. The UCST corresponds to the critical point at the top of the two-phase region, located at a critical composition \phi_c. This point satisfies the stability criteria \frac{\partial^2 \Delta G_{\text{mix}}}{\partial \phi^2} = 0 and \frac{\partial^3 \Delta G_{\text{mix}}}{\partial \phi^3} = 0, indicating an inflection in the free energy curve where the binodal and spinodal curves intersect and the distinction between phases vanishes. Entropy-enthalpy compensation underlies the temperature dependence of UCST, especially in systems driven by specific intermolecular interactions like hydrogen bonding. At low temperatures, forming a mixture requires breaking strong hydrogen bonds between like molecules to create weaker bonds between unlike pairs, yielding a positive \Delta H_{\text{mix}} that favors demixing. As temperature rises, the T \Delta S_{\text{mix}} term increases linearly, compensating and eventually surpassing the enthalpic penalty to drive dissolution above the UCST. These general principles extend to models such as Flory-Huggins theory for polymer solutions, where they predict phase behavior from interaction parameters.

Theoretical Models

Mean-Field Theory

Mean-field theory provides a foundational approximation for understanding phase separation in mixtures exhibiting an upper critical solution temperature (UCST), particularly through its application in regular solution theory. This approach treats the interactions between molecules as an average field experienced by each particle, effectively ignoring local fluctuations and correlations to simplify the statistical mechanics of the system. It assumes random mixing on a lattice where molecules of similar size occupy equivalent sites, making it suitable for binary mixtures of small, non-polar liquids without strong directional forces like hydrogen bonding. In regular solution theory, the key parameter is the Flory interaction parameter χ, which quantifies the non-ideal enthalpic contributions to mixing and is given by χ = ΔH_mix / (RT φ₁ φ₂), where ΔH_mix is the enthalpy of mixing, R is the gas constant, T is temperature, and φ₁, φ₂ are the volume fractions of the components. For symmetric mixtures (φ₁ = φ₂ = 0.5), phase separation occurs below the UCST when χ exceeds the critical value χ_c = 2, as this marks the point where the second derivative of the free energy of mixing with respect to composition vanishes, indicating instability of the homogeneous phase. The derivation stems from a lattice model where the solution is divided into sites, each with a coordination number z representing nearest neighbors. The enthalpic term arises from pairwise interactions: unlike A-A and B-B bonds contribute energies ε_{AA} and ε_{BB}, while A-B bonds contribute ε_{AB}, leading to an effective interaction energy Δε = ε_{AB} - (ε_{AA} + ε_{BB})/2. The UCST is then derived as the temperature where thermal energy balances these interactions at the critical point, yielding UCST = z Δε / (2R). Despite its simplicity, mean-field theory has limitations, as it neglects fluctuations that stabilize the disordered phase, resulting in an overprediction of the UCST compared to experimental values. This approximation is most accurate for small-molecule systems where molecular correlations are weak and the upper critical dimension is not exceeded. The model forms the basis for extensions to asymmetric systems, such as polymer solutions, via the .

Flory-Huggins Theory

The Flory-Huggins theory, developed independently by and in 1942, provides a mean-field lattice model for the thermodynamics of polymer solutions, accounting for the large size disparity between polymer chains and solvent molecules. The model derives the free energy of mixing from combinatorial entropy and an enthalpic interaction term, enabling predictions of phase behavior such as (UCST) in polymer-solvent systems. The χ parameter in this theory originates from mean-field approximations of pairwise interactions between polymer segments and solvent molecules. The entropy of mixing in the Flory-Huggins model incorporates a combinatorial factor that reflects the reduced configurational freedom of long polymer chains compared to solvent molecules. For a solution with polymer volume fraction φ and degree of polymerization N (number of segments per chain), the entropic contribution per lattice site to the Helmholtz free energy is given by \frac{\Delta F_s}{kT} = \frac{\phi}{N} \ln \phi + (1 - \phi) \ln (1 - \phi), where k is Boltzmann's constant and T is temperature; this expression arises from Stirling's approximation applied to the number of ways to arrange n_p = (total sites × φ / N) polymer chains and n_s = total sites × (1 - φ) solvent molecules on a lattice. For large N, the (φ / N) ln φ term becomes small, emphasizing the dominance of solvent entropy in dilute solutions. The full mixing free energy per site then includes an enthalpic term w φ (1 - φ), where w parameterizes segment-solvent interactions, leading to the dimensionless interaction parameter χ = w / (kT). In systems exhibiting UCST behavior, the interaction parameter χ displays temperature dependence of the form χ = A + B / T, where A and B are constants, with B > 0 reflecting enthalpically driven demixing upon cooling. occurs when χ exceeds a , given exactly by χ_c = \frac{1}{2} \left(1 + \frac{1}{\sqrt{N}}\right)^2 and approaching 0.5 in the limit of infinite N, derived from the stability condition at the spinodal (∂²(ΔG_mix)/∂φ² = 0) and intersection at the critical point. This threshold marks the θ-point where chains behave ideally in the infinite N limit. The theory predicts an asymmetrical for polymer solutions, with the curve skewed toward low polymer concentrations due to the chain connectivity and immobility, which limits polymer contributions compared to the . The UCST increases with increasing N, as larger chains require stronger interactions (higher χ at lower T) to overcome the reduced entropic penalty for demixing, resulting in narrower miscible regions for high-molecular-weight polymers. Extensions to the basic Flory-Huggins model incorporate specific interactions, such as hydrogen bonding, to better describe UCST in aqueous solutions where directional bonds between segments, , and possibly ions contribute to temperature-sensitive enthalpies. These modifications adjust the χ parameter to include bonding free energies, enabling predictions of closed-loop phase diagrams or salt-tolerant UCST behavior in systems like poly(N-acryloylglycinamide) in .

Experimental Examples

Binary Liquid Mixtures

Binary liquid mixtures of small molecules often exhibit upper critical solution temperature (UCST) behavior, where the components are fully miscible above the UCST but phase separate into two liquid phases below it due to dominant enthalpic repulsions that are overcome by thermal entropy at higher temperatures. A classic example is the hexane–nitrobenzene system, which has a UCST of 19 °C. In this mixture, the non-polar hexane and polar nitrobenzene show limited solubility at low temperatures owing to unfavorable intermolecular interactions, resulting in phase separation; above the UCST, increased kinetic energy promotes complete mixing. Another well-studied system is –water, with a UCST around 168 °C under sufficient to maintain . Here, phase immiscibility below the UCST arises from preferential hydrogen bonding within pure components— and —over mixed interactions, leading to segregation; elevated temperatures disrupt these bonds, enhancing through entropic contributions. The critical composition occurs at approximately 40 wt% . Phase diagrams for these symmetric UCST systems typically display a lens-shaped region of immiscibility, with the critical point near a 50:50 molar composition, as seen in the where the binodal curve meets at about 0.5 nitrobenzene. The consolute (UCST) temperature is commonly measured using the method, where a homogeneous is cooled until first appears, indicating the onset of . Additional examples include mixtures of perfluoroalkanes and s, such as –n-hexane, which exhibit UCST behavior due to weak interactions between the fluorinated and hydrocarbon components, resulting in at lower temperatures. These systems highlight the role of enthalpic drivers in UCST phenomena across diverse small-molecule binaries.

Polymer Solutions

solutions exhibit upper critical solution temperature (UCST) behavior when the mixture phase separates upon cooling below a critical , driven by enthalpic factors such as polymer-solvent incompatibilities or interchain interactions that dominate at lower temperatures. This contrasts with small-molecule systems by incorporating chain asymmetry and entanglement effects, leading to broader gaps and composition-dependent critical points. A classic example is the poly(ethylene oxide) (PEO)- system, where UCST occurs above 100°C; below this temperature, arises from strengthened hydrogen bonding between PEO chains, reducing . In this system, the interplay of hydrophilic groups and hydrophobic methylene segments results in a closed-loop , with restored at very high temperatures due to thermal disruption of ordered structures. The (PS)- system demonstrates UCST around 40°C, particularly in the semi-dilute regime, where cooling induces into polymer-rich and solvent-rich phases due to the theta-solvent nature of near . Experimental studies confirm this transition through coexistence curves, with the critical composition shifting toward lower polymer concentrations as molecular weight increases. In polymer-polymer blends, UCST behavior is evident in mixtures like polystyrene-poly(vinyl methyl ether) (PS-PVME), with a critical temperature around 140°C; phase separation upon cooling stems from weak intermolecular interactions and dynamic asymmetry between the rigid PS chains and flexible PVME segments. Such blends highlight how differing chain dynamics can stabilize or destabilize miscibility, often resulting in viscoelastic phase separation. UCST transitions in these polymer systems are commonly observed via light scattering techniques, which detect critical opalescence as intensity diverges near the critical point due to large fluctuations in composition. measurements, monitoring optical as a function of temperature, delineate the binodal curve by identifying the onset of macroscopic . Flory-Huggins theory briefly accounts for the observed in these systems through concentration-dependent interaction parameters.

Influencing Factors

Molecular and Compositional Effects

In polymer solutions exhibiting upper critical solution temperature (UCST) behavior, the UCST value increases with the molecular weight of the polymer, as longer chain lengths enhance intermolecular attractions, narrowing the miscible region below the UCST. This dependence arises from theoretical frameworks like Flory-Huggins theory, where the critical interaction parameter χ_c ≈ (1/2)(1 + 1/√N)^2 for chain length N, leading to a shift ΔT from the infinite molecular weight limit that scales approximately as ΔT / T_c ≈ 2 / √N for large N. For example, in polystyrene-based polyampholytes, increasing the from 20 to 97 raises the UCST from 46.5°C to 48.7°C in saline solutions. Concentration effects further modulate the UCST, with higher volume fractions φ elevating the temperature at which occurs, as increased chain density amplifies repulsive interactions above the critical point. In Flory-Huggins theory, the critical φ_c scales as approximately 1/√N for large N, indicating that dilute solutions of high-molecular-weight polymers exhibit broader but still shift UCST upward with rising φ. Experimental observations in polyampholyte solutions confirm this, showing UCST rising progressively from 1.0 to 5.0 g/L polymer concentration. In copolymers, the UCST can be precisely tuned by adjusting the comonomer composition to alter the hydrophilic-hydrophobic balance, enabling control over transitions in aqueous media. For instance, in -based copolymers with , increasing the fraction from 0.086 to 0.221 shifts the temperature from 5.5°C to 56.5°C at 1 mg/mL, as the more hydrophobic enhances polymer-polymer hydrogen bonding and aggregation below the UCST. This compositional tuning exploits the interplay between hydrophilic segments, which promote , and hydroneutral or hydrophobic comonomers that drive . For polymer-polymer blends, the UCST exhibits a strong dependence on blend ratios, often reaching a minimum at optimal compositions where specific interactions balance enthalpic and entropic contributions to miscibility. In blends like poly(ethylene oxide) with poly(methyl methacrylate-stat-styrene) copolymers, miscibility windows—regions of UCST behavior—expand or contract with varying styrene content, minimizing the UCST at compositions that optimize the Flory-Huggins interaction parameter χ across the blend. Such effects highlight how blend stoichiometry can stabilize homogeneous phases over wider temperature ranges.

Environmental Influences

External factors such as , ionic additives, and solvent composition can significantly modulate the upper critical solution temperature (UCST) in various systems, primarily by altering intermolecular interactions and thermodynamic stability of the mixed . typically elevates the UCST due to the negative volume change associated with demixing ( < 0), which favors the phase-separated state under compression according to the Clapeyron equation relating dependence to and changes. In polymer solutions like in trans-decalin, this manifests as an increase of approximately 5–6 °C per kbar. Some systems exhibit steeper shifts, up to 10–20 °C per kbar, highlighting the role of contractions upon mixing. The addition of salts influences UCST particularly in and zwitterionic solutions by screening electrostatic repulsions, thereby reducing the temperature required for . Ions follow the , where chaotropic ions more effectively lower the UCST through enhanced charge screening and altered water structure. For instance, in poly()-based systems, 1 M NaCl can depress the UCST by about 50 °C, promoting at lower temperatures. Similarly, in zwitterionic polyampholytes like P(VBTAC/NaSS), increasing NaCl concentration from 0 to 0.2 M reduces the UCST by over 20 °C at modest loadings. Co-solvents, such as , expand the miscible regime by improving quality and weakening polymer-polymer attractions, effectively shifting the UCST to lower values. Methanol or additions to aqueous or polymer solutions decrease cloud points, allowing broader temperature windows for homogeneity. This cosolvency effect is pronounced in systems like poly(oligo() methacrylate), where alcohol content linearly lowers the UCST while suppressing . UCST transitions often display temperature due to kinetic barriers, enabling below the equilibrium UCST without immediate . This arises from difficulties in forming the demixed phase, trapping the system in a metastable homogeneous . In solutions, such hysteresis can span several degrees , depending on cooling rates and solution .

Applications

Polymer Processing and Materials

In processing, the upper critical solution temperature (UCST) plays a pivotal role in enabling thermally induced (TIPS) for fabricating porous membranes without solvents, promoting environmentally friendly manufacturing. By cooling solutions below the UCST from a homogeneous achieved by heating above the UCST, demixing occurs, leading to solidification into hierarchical porous structures with controlled and high permeability. Analogous to classic UCST systems such as hexane-nitrobenzene mixtures, solutions can be processed via TIPS to form porous structures, where controlled cooling triggers fine demixing for enhanced membrane integrity and selectivity in industrial separation processes. UCST-driven temperature-controlled facilitates the development of sprayable formulations for coatings and adhesives, allowing polymers to remain dissolved at application temperatures and upon cooling to form adherent films. In coatings for building materials, UCST polymers enable precise deposition and improved by leveraging the sharp transition to create uniform layers without volatile organic compounds. For adhesives, this mechanism enhances bonding strength through rapid solidification at , suitable for industrial assembly lines. In polymer blend compatibilization, UCST behavior permits homogeneous mixing at elevated temperatures followed by controlled on cooling, which refines and interfacial properties. Reactive blends exhibiting UCST-type phase diagrams benefit from ionic functionalization—such as sulfonic acid-amine interactions—to form stable lamellar microphases, suppressing coalescence and yielding tougher plastics with enhanced impact resistance and mechanical integrity. Post-2020 advances have focused on tunable UCST polymers for resins, incorporating thermal reversibility to produce complex, responsive structures. Poly(acrylic acid-co-acrylamide) hydrogels, with UCST transitions around 25°C driven by hydrogen bonding, serve as for direct ink writing, enabling fabrication of flexible, hierarchical architectures for actuators that recover shape and properties upon temperature cycling. Polymer concentration influences these processing windows by shifting the UCST and modulating demixing rates.

Biomedical and Emerging Uses

In cell sheet engineering, UCST hydrogels such as those based on poly(N-acryloyl glycinamide-co-N-phenylacrylamide) enable non-invasive harvesting of adherent cells like NIH-3T3 fibroblasts by promoting cell attachment below the UCST (around 30°C) and spontaneous detachment upon mild heating to body temperature (37°C), preserving cell-cell junctions and integrity for applications. This thermoreversible behavior avoids enzymatic digestion, reducing cell damage compared to traditional methods. For , UCST-responsive micelles formed from peptide-mimetic polymers like PEG-b-P(NAGA-co-AN), where NAGA denotes N-acryloylglycinamide and AN , facilitate targeted release of therapeutics such as at physiological temperatures near 37°C, as the micelles disassemble above the UCST due to weakened hydrogen bonding, enabling complete payload discharge at tumor sites. These systems integrate with photothermal agents for on-demand activation, enhancing efficacy in cancer therapy while minimizing off-target effects. Thermoresponsive UCST polymers serve as sensors and actuators by exploiting phase transitions for dynamic responses; for instance, assemblies of PEG-b-(PEGMeA-co-PEGPhA)-b-PS exhibit nanoscale morphological changes upon heating above the UCST, enabling optical switches through reversible alterations in light scattering and properties. In soft robotics, UCST hydrogels like PNAGA-based bilayers actuate via bidirectional swelling-deswelling, providing mechanical deformation for biomimetic grippers or walkers in aqueous environments. Recent emerging applications (as of ) involve synergistic UCST-LCST polymers, such as PNAGA-PNIPAm hybrids, which respond to multiple stimuli (, , ) for advanced scaffolds that mimic dynamic extracellular matrices, supporting controlled and 3D tissue formation through tunable phase behaviors. Salt effects can fine-tune UCST transitions in these systems to optimize in physiological media.

References

  1. [1]
    IUPAC - upper critical solution temperature (UT07280)
    ### Summary of Upper Critical Solution Temperature (UCST)
  2. [2]
    Chapter 8 Phase Diagrams
    (b) Critical solution temperatures. The upper critical solution temperature (upper consolute temperature), Tuc, is the highest temperature at which phase ...
  3. [3]
  4. [4]
    How to manipulate the upper critical solution temperature (UCST)?
    Oct 21, 2016 · Polystyrene will dissolve in cyclohexane, in an upper critical solution temperature (UCST) type process. PNIPAM, on the other hand, will ...
  5. [5]
    [PDF] Dynamics and Microstructures of Nicotine/Water Binary Mixtures ...
    Sep 19, 2018 · completely miscible as the temperature is increased above the upper critical solution temperature (UCST). UCSTs are ubiquitous in binary ...
  6. [6]
    Polymers with Upper Critical Solution Temperature in Aqueous ...
    Jun 13, 2013 · Recently, much progress has been made in the synthesis of polymer systems that display UCST behavior under mild and physiologic conditions.
  7. [7]
    [PDF] Thermodynamics of mixing
    The highest temperature of the two-phase region is the upper critical solution temperature (UCST) Te. For all T> Tec, the homogeneous mixtures are stable ...
  8. [8]
    Equilibriums in systems of three components. Change of the mixing ...
    We then get the line de in which diudicates the critical mixing-temperature of phenol with water, and e that of aniline with water. ... critical solution of water ...Missing: upper consolute
  9. [9]
    A HISTORY OF SOLUTION THEORY - Annual Reviews
    Van Laar in 1906 delivered six lectures on "Thermodynamical Potential and Its Application to Problems in Chemical and Physical Equilibria." He began with the ...<|control11|><|separator|>
  10. [10]
    [PDF] Phase Diagrams and Phase Separation
    Spinodal decomposition occurs when any composition fluctuation is unstable – the unstable region of the phase diagram. Section of free energy curve: Page 17. AM ...<|control11|><|separator|>
  11. [11]
    [PDF] First principles thermodynamical modeling of the binodal and ...
    analyzed the consolute temperature or upper critical solution temperature (Tc) describing the lowest temperature at which both phases are miscible at any ...
  12. [12]
    [PDF] Mean Field Flory Huggins Lattice Theory
    Mean field: the interactions between molecules are assumed to be due to the interaction of a given molecule and an average field due to all the other.
  13. [13]
    [PDF] Mean Field Theory - Physics Courses
    ... mean field transition temperature. In the low temperature phase θ < 1, there are three solutions to the mean field equations. One solution is always at m ...
  14. [14]
    Thermodynamics of High Polymer Solutions - AIP Publishing
    Research Article| January 01 1942. Thermodynamics of High Polymer Solutions Available. Special Collection: JCP 90 for 90 Anniversary Collection. Paul J. Flory.Missing: original | Show results with:original
  15. [15]
    Phase Behavior of Polymer Solutions and Blends | Macromolecules
    We summarize our knowledge of the phase behavior of polymer solutions and blends using a unified approach. We begin with a derivation of the Flory–Huggins ...
  16. [16]
    Appendix B Flory–Huggins Theory of Polymer Solutions
    Flory parameter, a miscibility gap with an upper critical solution temperature (UCST) ... Such a behaviour is, for instance, observed in systems with hydrogen ...<|control11|><|separator|>
  17. [17]
    Constrained thermoresponsive polymers – new insights into ...
    Increasing interest is focussed on nonionic polymers in water, whose UCST behavior is based on hydrogen bonding. Due to their tolerance towards added salt ...
  18. [18]
    [PDF] Effect on the Upper Consolute Temperature (UCT) of a
    Jun 6, 2017 · It was found that the ionic compounds, which get hydrated with water, show lesser increase in CST as they decrease the miscibility to a lesser ...Missing: aniline | Show results with:aniline<|control11|><|separator|>
  19. [19]
    8.6: Phase Diagrams for Binary Mixtures - Chemistry LibreTexts
    Jul 19, 2021 · Similar behavior is seen for hexane/nitrobenzene mixtures, for which the critical temperature is 293 K. Figure 2.png Figure 8 . 6 . 2 :.
  20. [20]
    Reference Materials for Phase Equilibrium Studies. 1. Liquid ... - NIH
    Though the reported values of the upper consolute temperature range from 318 to 322 K, a relatively low value of 318.5 K from [27] has been selected because ...
  21. [21]
    [PDF] CHC 103 __ Solutions __ Partial miscibility of liquids_ Critical ...
    Therefore the nicotine water system shows an upper consolute temperature of 208 degrees celsius while lower consolute temperature of 61degrees Celsius. Thus ...
  22. [22]
    Cosolvency-Induced UCST/LCST Dual Phase Separation of Poly(2 ...
    Aug 26, 2025 · The cloud point for UCST-type and LCST-type phase separations was determined as the temperature at which the change in turbidity of the polymer ...
  23. [23]
    Liquid–liquid equilibrium of (perfluoroalkane + alkane) binary mixtures
    Apr 25, 2006 · As presented, the model could describe quite well the UCST of the mixtures but predicted narrower phase envelopes than the experimental ones.
  24. [24]
    Aqueous Solutions of Poly(ethylene oxide): Crossover from Ordinary ...
    The solubility of PEO in water has been studied for many years by cloud point measurements showing miscibility at low (LCST) and high temperatures (UCST) and ...
  25. [25]
    Upper-critical-solution-temperature behavior of the system ...
    Experimental temperatures range from 297 K to 306 K for the lower molar mass polymer in methylcyclohexane and 323 K to 331 K for the higher molar mass polymer ...Missing: history | Show results with:history
  26. [26]
    UCST behavior observed for a binary polymer mixture of polystyrene ...
    UCST behavior observed for a binary polymer mixture of polystyrene/poly(vinyl methyl ether) (PS/PVME) with a PS rich asymmetric composition as a result of ...
  27. [27]
  28. [28]
    The molecular weight dependence of lower and upper critical ...
    The structural term also has an effect on the molecular weight dependence of the familiar Upper Critical Solution Temperature. It is predicted that ...
  29. [29]
  30. [30]
  31. [31]
  32. [32]
  33. [33]
    Polymeric membranes made without organic solvent - ANR
    organic solvents by an original phase inversion method, i.e. temperature increase above the upper critical solution temperature of the polymeric system.
  34. [34]
    A review on thermally induced phase separation technology in the ...
    Aug 27, 2024 · ... upper critical solution temperature (UCST). The phase diagram demonstrates how the matrix of microporous polymer is highly achieved when the ...
  35. [35]
    FR3087446B1 - USE OF A UCST POLYMER IN A COATING ...
    The subject of the invention is the use, in a fluid aqueous composition capable of forming a surface coating of a support used in buildings and in the ...
  36. [36]
    Mussel-inspired hydrogels with UCST for temperature-controlled ...
    At the temperature higher than UCST (55 °C), all hydrogels show evidently enhanced adhesion strengths and PQAM-8 hydrogel exhibits the highest adhesion strength ...Missing: triggered | Show results with:triggered
  37. [37]
    Ionic Compatibilization of Polymers - ACS Publications
    Jul 22, 2022 · The corresponding unfunctionalized polymer blend demonstrated upper critical solution temperature (UCST)-type behavior. When the carboxylic ...Theoretical Considerations · Figure 4 · Figure 7
  38. [38]
    Multimaterial 3D printing of self-assembling smart thermo ...
    Jun 5, 2023 · In this review, we provide an overview of the functional design of the TRPs, the different classes of TRPs, and their application in 3D printing.
  39. [39]
    Temperature-Responsive Polymer Brush Coatings for Advanced ...
    Transitions governed by LCST (and UCST) of polymers result in a change in the thickness and wettability of polymer brush coatings [1,36].
  40. [40]
    Thermoresponsive Smart Copolymer Coatings Based on P(NIPAM ...
    Spontaneous cell detachment was reported for UCST-based materials composed of poly(N-acryloyl glycinamide-co-N-phenylacrylamide) copolymers, which were used as ...
  41. [41]
    Review Upper critical solution temperature polymeric drug carriers
    Mar 15, 2022 · Abstract. Stimuli-responsive polymers are widely used as drug carriers in therapeutic applications. Upper critical solution temperature (UCST) ...Missing: historical | Show results with:historical
  42. [42]
  43. [43]
    Upper critical solution temperature polymer assemblies via variable ...
    Jun 10, 2023 · For upper critical solution temperature (UCST) homopolymers, heating above the transition temperature results in an enthalpically driven ...
  44. [44]
    Synergistic Approaches in the Design and Applications of UCST ...
    Jul 21, 2023 · This review summarizes recent progress in the synergistic design strategy for thermoresponsive polymers possessing an upper critical solution temperature (UCST ...