Fact-checked by Grok 2 weeks ago

Lower critical solution temperature

The lower critical solution temperature (LCST) is the critical temperature below which the components of a binary mixture are completely miscible in a single homogeneous phase and above which the mixture undergoes phase separation into two immiscible liquid phases. This is in contrast to the upper critical solution temperature (UCST), above which the mixture is miscible and below which it separates. This phenomenon is observed in various liquid-liquid systems, including aqueous solutions of ionic liquids, non-polar polymer-solvent mixtures, and certain small-molecule binaries, where strong intermolecular interactions like hydrogen bonding promote miscibility at lower temperatures. Thermodynamically, LCST behavior arises from the Gibbs free energy of mixing, \Delta G_\text{mix} = \Delta H_\text{mix} - T \Delta S_\text{mix}, where a negative enthalpy of mixing (\Delta H_\text{mix} < 0) due to favorable interactions dominates at low temperatures, but as temperature increases, the entropic term (-T \Delta S_\text{mix}) becomes more positive, driving phase separation when like-like interactions (e.g., hydrophobic effects) outweigh unlike interactions. In polymer systems, such as poly(N-isopropylacrylamide) (PNIPAM) in water, the LCST typically occurs around 32°C and is influenced by factors like chain length, concentration, and specific interactions, often modeled using extensions of Flory-Huggins theory that incorporate temperature-dependent interaction parameters. LCST materials are notable for their thermoresponsive properties, enabling applications in drug delivery, smart hydrogels, and separation processes like desalination, where heating induces reversible phase changes without additional chemicals.

Introduction

Definition

The lower critical solution temperature (LCST), also known as the lower consolute temperature, is the critical temperature below which the components of a mixture of two or more substances are completely miscible in all proportions under given pressure conditions. This phenomenon is observed in certain binary or multicomponent liquid mixtures where solubility increases at lower temperatures and decreases upon heating. The LCST marks the point where the mixture transitions from a single homogeneous phase to partial immiscibility, with the two coexisting phases becoming indistinguishable exactly at the critical point. In mixtures exhibiting an LCST, the system remains fully miscible and forms a single phase below the LCST, while above this temperature, phase separation occurs into two immiscible liquid phases of differing compositions. The miscibility gap—the region of partial miscibility—originates at the LCST and widens as temperature rises, driven by changes in intermolecular interactions that favor demixing at higher thermal energies. The critical point at the LCST defines the boundary of this gap, where the compositions of the two phases converge, and properties such as density, refractive index, and interfacial tension approach zero difference between phases. On a temperature-composition phase diagram, the LCST appears as the lower vertex of a lens-shaped two-phase region that opens upward with increasing temperature. The binodal curve delineates the boundaries of the miscible and immiscible regions, enclosing the area where phase separation occurs for compositions between the critical point and the pure-component limits. This upward-opening configuration contrasts with (UCST) behavior and highlights the temperature-dependent solubility reversal characteristic of LCST systems. The LCST phenomenon was observed in 1898 by V. Rothmund in the triethylamine-water system, where the mixture shows complete miscibility below approximately 18–19 °C and separates into two phases upon slight warming. This early discovery laid the foundation for understanding consolute temperatures in liquid mixtures, with the LCST representing one end of the spectrum of critical solution behaviors.

Comparison with Upper Critical Solution Temperature

The upper critical solution temperature (UCST) is defined as the critical temperature above which a mixture of two or more components is completely miscible in all proportions, with phase separation occurring below this temperature primarily due to enthalpic dominance that favors immiscibility at lower temperatures. In UCST systems, the miscibility increases with temperature as thermal energy overcomes unfavorable enthalpic interactions, such as those from van der Waals forces, leading to a phase diagram with a miscibility gap below the critical point. In comparison, the lower critical solution temperature (LCST) marks the point below which complete miscibility occurs, but above which phase separation takes place due to entropy-driven effects, such as the disruption of structured hydrogen bonding or hydration shells that reduces configurational entropy upon heating. This fundamental difference arises from the thermodynamic drivers: UCST phase behavior is enthalpically controlled, with positive entropy of mixing, whereas LCST is entropically disfavored at higher temperatures, often accompanied by negative enthalpy of mixing. LCST is typically observed in aqueous or polar associating systems, while UCST predominates in non-polar solvent environments. Certain mixtures, particularly polymer blends, can exhibit both UCST and LCST behaviors, forming rare closed-loop miscibility gaps in their phase diagrams where a narrow temperature range of complete miscibility is bounded by upper and lower critical curves. From a practical standpoint, LCST properties facilitate the design of temperature-responsive materials, such as those used in drug delivery and tissue engineering, enabling reversible phase transitions near physiological temperatures. In contrast, UCST behaviors are leveraged for solvent-resistant polymers and membranes, where phase separation below the critical temperature supports applications requiring stability in organic solvents.

Physical Basis

Thermodynamic Principles

The phase behavior associated with the (LCST) is governed by the Gibbs free energy of mixing, \Delta G_\text{mix}, which determines the stability of homogeneous mixtures versus phase-separated states. The expression \Delta G_\text{mix} = \Delta H_\text{mix} - T \Delta S_\text{mix} encapsulates the competition between enthalpic and entropic contributions, where below the LCST, the free energy minimum favors a single mixed phase, while above it, the system separates into two coexisting phases to lower the overall free energy. At the LCST, the unfavorable entropy term begins to dominate due to ordered structures in the low-temperature mixed phase, shifting the free energy landscape to promote demixing upon heating. Enthalpy plays a key role in stabilizing the mixed state at low temperatures, with negative \Delta H_\text{mix} arising from attractive interactions that favor miscibility. For instance, exothermic mixing enthalpies overcome the entropic penalty from ordered configurations, such as clustered or structured arrangements in the solution. Upon heating, the entropy change \Delta S_\text{mix} for the mixing process is negative, reflecting the higher entropy of the demixed state where order is disrupted; the -T \Delta S_\text{mix} term thus grows in magnitude, driving phase separation as the entropic favorability of separation outweighs the enthalpic gain. This counterintuitive temperature dependence—where heating reduces solubility—highlights how LCST systems exhibit enthalpy-driven mixing at low temperatures and entropy-driven demixing at higher ones. The LCST marks the critical point where the homogeneous phase becomes unstable, defined by the conditions \left( \frac{\partial^2 \Delta G_\text{mix}}{\partial \phi^2} \right)_{T,P} = 0 and \left( \frac{\partial^3 \Delta G_\text{mix}}{\partial \phi^3} \right)_{T,P} = 0 at the critical composition \phi_c, with \phi denoting the volume fraction. These criteria indicate an inflection point in the free energy curve, where the second derivative vanishes (spinodal condition) and the third ensures the critical nature, bounding the miscibility gap below which the mixture is stable against infinitesimal composition fluctuations.

Molecular Interactions

The lower critical solution temperature (LCST) behavior arises primarily from associative intermolecular interactions, such as hydrogen bonding and dipole-dipole forces, which stabilize miscible phases at low temperatures but weaken upon heating. At lower temperatures, these interactions promote the formation of ordered networks between solute molecules (e.g., polymer chains) and solvent molecules, resulting in a negative enthalpy of mixing that favors solubility. Upon increasing temperature, thermal energy disrupts these bonds, leading to a positive enthalpy change and reduced solubility, thereby driving phase separation. This temperature-dependent breakage of associative forces is a key molecular driver of LCST, as evidenced in systems like poly(N-isopropylacrylamide) (PNIPAM) where hydrogen bonds between amide groups and water dominate below the transition temperature. In aqueous systems, the hydrophobic effect further contributes to LCST by imposing an entropy penalty on solvation of non-polar groups at lower temperatures. Water molecules form structured cages around hydrophobic moieties, reducing the overall entropy of the system and stabilizing the dissolved state through enthalpic compensation from hydrogen bonding within the solvent network. As temperature rises, this structuring diminishes, releasing the entropy penalty and making hydrophobic associations between solute molecules more favorable, which promotes aggregation and demixing. This entropic relief is particularly pronounced in amphiphilic polymers, where the balance between hydrophilic and hydrophobic segments dictates the sharpness of the LCST transition. For polymeric systems, the flexibility of polymer chains influences LCST through changes in conformational entropy. At low temperatures, extended chain conformations maximize entropy in the solvated state, supported by favorable solvent-polymer interactions. Heating induces chain collapse, reducing conformational entropy but outweighed by gains in solvent entropy from disrupted hydration shells, resulting in net phase separation. This conformational shift is entropy-driven and amplifies the hydrophobic effect in flexible chains like those in PNIPAM. Additives such as cosolvents and salts modulate these molecular interactions, often via salting-out effects that alter values. Kosmotropic salts (e.g., those following the like ) compete for water molecules, weakening polymer hydration and effectively enhancing hydrophobic interactions, which typically lowers the LCST by facilitating earlier phase separation. In contrast, certain cosolvents can either stabilize or disrupt hydrogen bonding networks; for instance, polar cosolvents may increase LCST by improving solvation entropy, while amphiphilic ones promote collapse through preferential hydrophobic solvation. These modulations highlight how external species fine-tune the balance of associative and hydrophobic forces underlying LCST behavior.

Systems Exhibiting LCST

Polymer-Solvent Mixtures

In polymer-solvent mixtures, the lower critical solution temperature (LCST) behavior is influenced by the high molecular weight of polymers, which results in broader miscibility gaps compared to low-molecular-weight systems due to diminished entropy of mixing and stronger enthalpic contributions from chain interactions. This leads to wider regions of phase immiscibility in temperature-concentration phase diagrams, where solutions separate into polymer-rich and solvent-rich phases above the LCST. The LCST itself often decreases with increasing polymer molecular weight, as longer chains enhance hydrophobic associations that drive phase separation at lower temperatures. A prototypical example of LCST in polymer-solvent mixtures is poly(N-isopropylacrylamide) (PNIPAM) dissolved in water, where the homopolymer exhibits an LCST of approximately 32°C, marking the onset of coil-to-globule transition and phase separation. This transition is highly sensitive to environmental factors: the LCST decreases with added salts (particularly kosmotropic anions that weaken hydration shells) or higher molecular weight, while it increases under basic pH conditions due to altered amide ionization. PNIPAM's responsiveness stems from temperature-dependent hydrogen bonding between its amide groups and water molecules, which weakens above the LCST, favoring intra- and intermolecular hydrophobic interactions. Other notable polymers displaying LCST include poly(vinyl methyl ether) (PVME) in water, with a transition around 37°C driven by ether-water hydrogen bonds that rupture at higher temperatures, and derivatives of poly(ethylene glycol) (PEG) such as poly(oligo(ethylene glycol) methacrylate) (POEGMA), which show LCST values tunable from 26°C to over 90°C depending on side-chain length. Copolymerization significantly shifts the LCST; for instance, incorporating hydrophilic monomers like N,N-dimethylacrylamide into PNIPAM raises the LCST by enhancing water solubility, whereas hydrophobic comonomers like n-butyl methacrylate lower it by promoting aggregation. Grafting architectures, such as PNIPAM brushes on surfaces, further tune the LCST upward with increased grafting density, as constrained chains reduce conformational entropy gains upon dissolution. Experimentally, LCST in these mixtures is observed through cloud point measurements, where the temperature at which the solution becomes turbid—indicating macroscopic phase separation—is determined by monitoring transmittance via UV-Vis spectrophotometry or dynamic light scattering to track aggregate formation. These techniques provide the cloud point temperature (T_cp), often taken as the inflection point of turbidity curves, offering a practical proxy for the LCST in dilute solutions.

Small Molecule Mixtures

Small molecule mixtures exhibiting (LCST) behavior demonstrate phase separation upon heating, typically in binary liquid systems where associating components like hydrogen-bonding or electrostatic interactions dominate. Unlike polymeric systems, these mixtures often display faster dynamics due to the absence of chain entanglements and narrower miscibility gaps, allowing for rapid equilibration and reversible transitions driven by temperature-dependent molecular associations. A classic example is the triethylamine-water binary mixture, which shows an LCST of 18.3 °C at a critical composition near 32 mass% triethylamine, where the two liquids are fully miscible below this temperature but phase separate into conjugate solutions upon slight heating. This behavior arises from weakened hydrogen bonding between water and the amine groups at higher temperatures, leading to clustering of triethylamine molecules. Similarly, the nicotine-water system exhibits an LCST around 60 °C, with orientational dynamics slowing anomalously near the critical concentration due to persistent hydrogen-bonded structures between nicotine's pyridine moiety and water, as observed via optical heterodyne-detected optical Kerr-effect spectroscopy and NMR. In non-aqueous systems, perfluorocarbons mixed with hydrocarbons provide another illustration, where the lipophobic nature of perfluorocarbons results in LCSTs often below room temperature; for instance, perfluorooctyl bromide in n-hexane has an LCST approximately 40 °C lower than perfluorodecalin in the same solvent, reflecting differences in intermolecular cohesivity and solubility. Ionic liquids also show LCST in binary mixtures, such as 1,3-dimethylimidazolium iodide in acetone, with cloud points tunable from 44.8 °C to 49.4 °C depending on concentration (300–500 mg mL⁻¹), controlled by cation-anion affinity and solvent choice rather than hydrogen bonding. Choline-based ionic liquids, like N-alkyl-N,N-dimethylhydroxyethylammonium bis(trifluoromethane)sulfonylimide with ethers, exhibit unusual LCST-type demixing between 200 K and 360 K, influenced by alkyl chain length and ether dipole exposure. Recent studies have expanded LCST applications in energy-related systems, particularly water-ionic liquid mixtures for CO₂ capture. Polyetheramine-fatty acid ionic liquids in water display reversible LCST transitions from 37 °C to 91 °C, with dual responsiveness to temperature and CO₂: bubbling CO₂ induces phase separation via bicarbonate formation, while heating enhances droplet aggregation from 6.5 nm to 21.0 nm, enabling efficient capture and release cycles. Small-molecule 1,2,3-triazolium ionic liquids with tailored hydrophobicity, such as [Bu-i-Hex-C2OH-tr][OTMBS] (LCST 17 °C), further demonstrate tunable phase behavior in water, with critical temperatures adjustable from 5 °C to 56 °C by varying alkyl chains, highlighting their potential in thermoresponsive separations.

Theoretical Frameworks

Flory-Huggins Theory

The Flory-Huggins theory provides a foundational lattice-based mean-field framework for describing the thermodynamics of polymer solutions, particularly in understanding phase behavior such as the (LCST). Developed independently by and in 1942, the model represents the solution as a regular lattice where solvent molecules and polymer segments occupy sites, accounting for the large size disparity between solvent and polymer chains. The core of the theory is the expression for the Gibbs free energy of mixing, \Delta G_\text{mix}, given by \Delta G_\text{mix} = RT \left[ n_1 \ln \phi_1 + n_2 \ln \phi_2 + \chi n_1 \phi_2 \right], where R is the gas constant, T is the temperature, n_1 and n_2 are the numbers of moles of solvent and polymer, \phi_1 and \phi_2 are the volume fractions of solvent and polymer (\phi_1 + \phi_2 = 1), and \chi is the Flory interaction parameter that captures the enthalpic interactions between unlike species relative to like species. This formulation arises from combining the entropic contribution due to random mixing on the lattice with an enthalpic term proportional to the product of volume fractions of the components. The interaction parameter \chi is temperature-dependent and central to predicting phase separation. In its simplest form, \chi \propto 1/T, but more generally, it is expressed as \chi = A + B/T, where A and B are constants reflecting entropic and enthalpic contributions, respectively. For upper critical solution temperature (UCST) behavior, B > 0, causing \chi to increase as temperature decreases, leading to phase separation upon cooling when \chi exceeds a critical value. For LCST behavior, B < 0 (or equivalently \chi = A - B/T with B > 0), resulting in \chi increasing with rising temperature due to an entropy-driven mechanism, often linked to specific interactions like hydrogen bonding that weaken at higher temperatures; this predicts miscibility at low temperatures and demixing above the LCST when \chi(T) = \chi_c. Phase separation in polymer-solvent mixtures is predicted when \chi > \chi_c, with the critical interaction parameter given by \chi_c = 0.5 (1 + 1/\sqrt{N}), where N is the of the . For large N, \chi_c \approx 0.5, and the LCST occurs at the where \chi(T) reaches this value, marking the onset of immiscibility; the critical polymer volume fraction is approximately $1/\sqrt{N}. This allows estimation of LCST from measured \chi(T) and polymer chain length, as demonstrated in systems like poly()-water mixtures. Despite its utility, the Flory-Huggins theory has limitations, including the assumption of incompressibility (fixed lattice volume, neglecting free volume differences), mean-field averaging that ignores concentration fluctuations and chain connectivity effects, and inability to capture strong specific associations like hydrogen bonding without modifications; these shortcomings particularly affect accurate LCST predictions in real systems with orientational order or volume changes upon mixing.

Advanced Models

Building on the foundational Flory-Huggins framework, advanced models for lower critical solution temperature (LCST) address limitations such as incompressibility assumptions by incorporating more realistic physical effects like volume changes and specific interactions. The Sanchez-Lacombe lattice fluid theory represents a key extension, treating mixtures as compressible fluids through an equation-of-state approach that explicitly accounts for free volume effects. This model calculates the of mixing by considering the occupancy of sites by molecules and vacant sites (holes), allowing for density fluctuations and compressibility that drive LCST in polymer solutions. By deriving thermodynamic properties like chemical potentials from the equation of state, it predicts LCST values semi-quantitatively for a wide range of polymer-solvent systems, such as in , where free volume disparities become prominent at higher temperatures. Associating models, such as those based on Wertheim's or perturbed hard-sphere chain approaches, enhance predictions for systems involving ing by decomposing the Flory-Huggins interaction parameter χ into enthalpic, , and associative contributions. Wertheim's theory models association as a resummed perturbation on a reference hard-sphere fluid, capturing the formation and breakage of s that lead to temperature-dependent χ, often resulting in LCST behavior in aqueous solutions like poly(N-isopropylacrylamide). In these frameworks, the associative term reflects the loss from networks, which strengthens at low temperatures to promote but weakens upon heating, inducing . This approach has been applied to continuum-space representations of brushes, accurately reproducing LCST transitions by balancing and bonding effects. For mixtures, adaptations of regular solution theory incorporate solvophobic interactions to explain rare LCST phenomena, where arises from rather than purely enthalpic drivers. Traditional regular solution theory assumes a temperature-independent χ based on parameters, but extensions introduce a solvophobic potential—modeled as a repulsive interaction that increases with —to account for structured around nonpolar solutes, leading to aggregation above the LCST. This is evident in systems like ionic liquid-water mixtures, where solvophobic aggregation of hydrophobic moieties disrupts shells, mimicking polymer-like LCST behavior without chain dominance. Such modifications enable prediction of LCST in non-aqueous blends, emphasizing the role of formation and solvent reorganization. Recent advances integrate with thermodynamic models to capture associative in LCST predictions, particularly for complex systems post-2020. frameworks, such as multitask neural networks trained on experimental χ parameters and solubility data, predict temperature-dependent by learning hidden entropic contributions from and hydrophobicity descriptors like values. These models achieve R² ≈ 0.83 for Flory-Huggins χ parameter predictions, while sparse modeling approaches using exhaustive search on small datasets forecast temperatures with RMSE ≈ 5-8 °C for thermo-responsive copolymers, indirectly modeling associative through correlations with and concentration, enabling rapid screening of new materials without full equation-of-state simulations. For instance, sparse modeling approaches using exhaustive search on small datasets forecast LCST shifts due to compositional changes, highlighting entropic penalties from associative networks in aqueous environments. Additionally, as of 2025, self-driving laboratories employing have been used to autonomously optimize LCST in poly(N-isopropylacrylamide) solutions with salts, accelerating experimental .

Prediction Methods

Empirical Approaches

Empirical approaches to determining the lower critical solution temperature (LCST) rely on laboratory-based experimental techniques that identify the onset of , often through observable changes in mixture clarity or thermal properties. These methods provide direct measurements of or transition temperatures, forming the basis for subsequent correlative analyses. One foundational technique is visual observation, where the temperature at which a transparent turns cloudy—marking the —is recorded as the LCST indicator. This method, suitable for low-concentration mixtures, has been applied to polymer-solvent systems like poly(2-chloroethyl vinyl ether) in water, allowing simple detection without specialized equipment. Early 20th-century studies on liquid mixtures employed similar visual assessments, often involving incremental heating or cooling to map boundaries, though lacking modern precision. For more quantitative detection, nephelometry and related turbidity measurements monitor light scattering or intensity as temperature rises, capturing the sharp increase in opacity at the LCST. In aqueous poly(N-isopropylacrylamide) solutions, UV has been used to precisely determine cloud points across concentrations from 0.5 to 22 wt%, yielding results comparable to other methods under conditions. These optical techniques are particularly effective for thermoresponsive polymers, providing real-time data on phase transitions. Differential scanning calorimetry (DSC) offers a thermal perspective by detecting endothermic peaks associated with the LCST phase separation, quantifying changes and transition temperatures. High-sensitivity DSC applied to poly(ethylene oxide) in revealed LCST values decreasing with concentration (e.g., from ~130°C at low concentrations), alongside positive increments indicative of hydrophobic exposure during the transition. This method complements optical approaches by elucidating the energetics of mixing. Experimental LCST data are frequently analyzed using empirical correlations to relate transition temperatures to mixture composition or molecular features. Linear relationships between the inverse LCST (1/T_LCST) and have been observed in systems, enabling prediction of across concentration ranges. Several factors measured LCST values, requiring careful in experiments. Molecular weight effects vary by ; for poly(N-isopropylacrylamide), end-group dominates at low weights, but higher weights generally lower the LCST due to reduced end-group . Pressure alters the , as seen in poly(vinyl methyl ether)- mixtures where elevated pressure narrows the LCST region across compositions. Additives like salts typically depress the LCST through disruption of hydration shells, as demonstrated in poly(N-isopropylacrylamide) solutions where reduces temperatures.

Computational Predictions

Quantitative structure-property relationship (QSPR) models predict the lower critical solution temperature (LCST), often denoted as θ, by correlating molecular descriptors of polymers and solvents with experimental θ values through techniques. These descriptors include topological indices that capture molecular and quantum chemical properties such as distributions, enabling the estimation of θ without direct experimentation. For instance, a multivariate model using norm-based topological and quantum descriptors achieved a (R²) of 0.942 and a mean relative (MRE) of 2.65% on a of 118 polymer-solvent systems. Such models facilitate rapid screening of candidate materials for LCST behavior by inputting structural data into predefined equations. Molecular dynamics (MD) simulations provide atomistic insights into LCST transitions by modeling polymer-solvent interactions over time to compute the temperature-dependent Flory-Huggins interaction parameter χ(T). All-atom MD employs detailed force fields to simulate molecular trajectories, revealing how hydrogen bonding and hydrophobic effects drive phase separation as temperature increases. Coarse-grained approaches simplify this by representing groups of atoms as beads, allowing longer simulations to locate critical points where χ(T) reaches values indicative of demixing, such as χ ≈ 0.5. For example, simulations of ionic liquid-water mixtures used radial distribution functions from MD trajectories to identify LCST at approximately 36°C, correlating peak shifts in ion-water correlations with the onset of clustering. These methods predict LCST for specific systems like poly(N-isopropylacrylamide) in aqueous environments by analyzing density fluctuations and solubility trends across temperature ramps. Recent coarse-grained models, such as Mpipi-T developed in 2025, enable prediction of LCST-type phase transitions in intrinsically disordered proteins by incorporating atomistic solvation free energies and temperature-dependent interactions. The theta temperature θ serves as the LCST analog at infinite polymer dilution, defined as the point where the second virial coefficient vanishes and polymer chains exhibit ideal coil behavior in . At θ, the polymer-solvent parameter χ equals 0.5, balancing enthalpic and entropic contributions to . Predictions of θ leverage parameters δ of the (δ_p) and (δ_s), with the χ ≈ (V/RT)(δ_p - δ_s)^2 linking structural to strength, where V is the , R the , and T the temperature. This approximation allows estimation of θ by solving for the temperature where |δ_p - δ_s| ≈ √(0.5 RT / V), using tabulated or computed δ values to forecast miscibility limits. Group contribution methods estimate these parameters by aggregating contributions from molecular groups. Recent advancements in and (AI/ML) have enhanced LCST predictions by integrating quantum descriptors and polymer fingerprints into data-driven frameworks, achieving accuracies exceeding 90% on diverse datasets from 2021 to 2025. Support vector regression models, for instance, predicted LCST shifts in poly(N-isopropylacrylamide) copolymers under varying salt conditions with high fidelity, using features like and copolymer composition. A 2023 polymer fingerprint approach, combining SMILES-based representations with artificial neural networks, yielded an R² of 0.95 and root mean squared error of 5.3°C across 893 polymer entries, enabling to novel thermoresponsive materials. These AI/ML tools, often incorporating advanced theoretical models as feature inputs, support high-throughput design of LCST systems for applications like .

Applications and Recent Developments

Biomedical Applications

Temperature-responsive hydrogels based on polymers exhibiting lower critical solution (LCST) behavior have emerged as versatile platforms in biomedical applications, leveraging their ability to undergo reversible phase transitions near physiological s for precise control over material properties. These hydrogels, often composed of poly(N-isopropylacrylamide) (PNIPAM) or its copolymers, swell in an extended, hydrophilic state below the LCST, facilitating encapsulation and retention of therapeutic agents, and collapse into a compact, hydrophobic form above the LCST, enabling triggered release. This thermoresponsive mechanism aligns well with body (approximately 37°C), making such systems biocompatible and suitable for in vivo use without external stimuli in many cases. In , PNIPAM-based hydrogels exemplify LCST exploitation for controlled release, where the polymer's LCST (typically 32–33°C, tunable via copolymerization) allows sustained payload delivery at targeted sites. For instance, injectable PNIPAM hydrogels loaded with demonstrate temperature-triggered deswelling at 37°C, enabling controlled, temperature-triggered release while minimizing burst effects and in cancer models. Similarly, hybrid systems incorporating or enhance biocompatibility and mechanical strength, supporting applications like through vascular endothelial growth factor release upon mild . These properties reduce dosing frequency and improve therapeutic efficacy compared to non-responsive carriers. For and , LCST polymers enable smart surfaces that support reversible and non-invasive harvesting, critical for . PNIPAM-grafted substrates promote attachment and proliferation at 37°C (above LCST, hydrophobic state favoring ) and allow intact sheet detachment by cooling to 20–25°C (below LCST, hydrophilic swelling disrupts interactions). This approach, pioneered in cell sheet engineering, has facilitated corneal tissue reconstruction in clinical trials, where multilayered epithelial sheets integrate seamlessly without sutures. In tissue scaffolds, LCST hydrogels mimic dynamics, enhancing chondrogenesis in repair models through tunable and nutrient . LCST-driven phase changes also underpin biosensors and diagnostics, particularly for protein purification and analyte detection. Elastin-like polypeptides (ELPs), recombinant proteins with tunable LCST (20–40°C), enable inverse transition cycling for non-chromatographic purification: fusion proteins remain soluble below LCST, bind targets, and aggregate above LCST for facile separation, achieving >95% purity in one cycle with recyclability. In biosensing, PNIPAM-modified microarrays detect glucose or steroids via temperature-modulated swelling, which alters analyte accessibility and signal output, offering sensitivity down to micromolar levels. These systems provide cost-effective alternatives to traditional methods, with applications in point-of-care diagnostics. Recent advances in the have focused on biocompatible LCST polymers for injectable therapeutics, addressing challenges like gelation and biodegradability. For example, glycol chitin-based hydrogels exhibit sol-to-gel at 37°C, enabling minimally invasive of antimicrobials for biofilm disruption in chronic infections, with degradation profiles matching tissue remodeling. Copolymer systems like PNIPAM blended with (PEG) have advanced ocular injectables, forming temporary seals post-surgery in animal models while ensuring complete resorption. These developments, including ELP fusions for targeted protein therapeutics, emphasize enhanced and reduced , paving the way for clinical translation in . As of , LCST properties have been explored in shape memory polymers for and gels for advanced , further expanding clinical potential.

Industrial and Emerging Uses

Lower critical solution temperature (LCST) phenomena enable efficient solvent recycling in processing, particularly in continuous of propylene-based polymers. In such processes, a gravimetric exploits the LCST boundary to separate polymer-rich and polymer-lean after , allowing the lean phase—containing unreacted and solvents—to be heated (typically to 140–220°C) and recycled back to the without issues or the need for energy-intensive compressors. This approach enhances plant capacity by supporting higher concentrations and reduces operational costs through streamlined recycle streams maintained above the polymer's temperature to prevent solidification. LCST properties also facilitate sustainable fiber spinning via environmentally friendly techniques. Thermoresponsive polymers, such as poly(N-isopropylacrylamide), dissolve in water below their LCST (around 32°C), enabling solvent-free or low-solvent aqueous spinning under controlled humidity and to form nanofibers without toxic solvents like or DMF. Post-spinning crosslinking stabilizes the fibers while preserving thermoresponsiveness, minimizing environmental residues and health risks associated with traditional methods. In separation technologies, LCST-driven aqueous two-phase systems (ATPS) provide recyclable platforms for purifying biomolecules in pharmaceutical and . Thermo-responsive copolymers like poly(N-vinylbutyramide-co-acrylamide) exhibit LCST transitions that allow and (up to 95% efficiency) after , enabling in ATPS for isolating active pharmaceutical ingredients or proteins without volatile solvents. Similar systems support food-grade separations, such as from broths, by leveraging the and mild conditions of water-based phases. Emerging applications harness LCST for energy storage in phase-change systems integrated with water management. LCST mixtures, such as those involving organic acids or salts, undergo thermally induced phase separation to store latent heat in processes like desalination or atmospheric water harvesting, where the partial molar enthalpy of water in the dilute phase must be sufficiently negative (e.g., 2.5 times more than current benchmarks) for efficient energy recovery. Recent analyses indicate that optimizing chemical potentials could enhance refrigeration and dehumidification cycles, though higher heat inputs are required for improved mixtures. Post-2021 research highlights LCST ionic liquids for CO2 capture, leveraging temperature-triggered transitions for reversible . Small-molecule thermoresponsive ionic liquids, like amino-functionalized imidazolium variants, achieve up to 49.6% CO2 at by forming single s below LCST for uptake and separating above it for release, outperforming non-responsive analogs by enabling energy-efficient regeneration. Studies from 2022–2023 emphasize nanostructured IL-water mixtures that enhance selectivity and reduce energy penalties in capture processes. Sustainability efforts in utilize LCST to minimize volatile organic compounds through water-based systems. Biorenewable solvents like are upgraded via copper-mediated into water-soluble LCST poly(lactamide acrylates), yielding tunable cloud points (12–62°C) for thermo-responsive materials without petroleum-derived feedstocks. By 2024–2025, such approaches have advanced reversible salting-out in aqueous LCST formulations for sustainable and processing, reducing solvent waste and promoting circular polymer economies.

References

  1. [1]
    Lower Critical Solution Temperature Phase Behavior and Water ...
    Oct 7, 2025 · Mixtures that possess a lower critical solution temperature (LCST) phase behavior form a homogeneous single phase at temperatures below the LCST ...
  2. [2]
    Molecular insight into the lower critical solution temperature ... - Nature
    May 2, 2019 · In liquid–liquid mixtures with a LCST transition, a single and miscible phase appears at lower temperatures. However, upon heating above a ...
  3. [3]
    [PDF] Mean Field Flory Huggins Lattice Theory
    To aid in modeling, the solution is imagined to be divided into a set of cells within which molecules or parts of molecules can be placed (lattice theory). ...
  4. [4]
    lower critical solution temperature (12244) - IUPAC Gold Book
    Critical temperature below which a mixture is miscible at any composition. Notes: The LCST depends upon pressure and the molar-mass distribution functions of ...
  5. [5]
    Lower Critical Solution Temperature - ScienceDirect.com
    Lower Critical Solution Temperature (LCST) is the temperature where a substance becomes insoluble in a solvent, leading to precipitation when the temperature ...<|control11|><|separator|>
  6. [6]
    Thermodynamic properties of lower critical solution temperature ...
    Jun 25, 2025 · Mixtures that possess a lower critical solution temperature (LCST) phase behavior are homogeneous at temperatures below the LCST and separate ...
  7. [7]
    IUPAC-NIST Solubility Data Series. 96. Amines with Water Part 1. C 4
    Dec 19, 2012 · Rothmund, Z. Phys. Chem., Stoechiom. Verwandtschaftsl. 26, 433 (1898). (2) Water; H2O; [7732-18-5]. Variables: Prepared by: Temperature ...
  8. [8]
  9. [9]
  10. [10]
  11. [11]
    Thermoresponsive polymers with lower critical solution temperature
    Thermoresponsive polymers with lower critical solution temperature (LCST) undergo reversible phase transitions and are used as smart materials. This article ...
  12. [12]
  13. [13]
  14. [14]
  15. [15]
    [PDF] Thermodynamics of mixing
    (4.31), then x decreases as temperature is lowered. The lowest tem- perature of the two-phase region is the lower critical solution temperature. (LCST), and ...
  16. [16]
    The role of hydrogen bonding in tethered polymer layers - PMC - NIH
    Namely, that the solubility of PEO decreases with increasing temperature and then the solution shows a lower critical temperature for phase separation26-38.
  17. [17]
    Understanding the lower critical solution temperature of amphiphilic ...
    Jul 1, 2025 · Considering first the hydrogen bonding of water to polymer, simulations detect a lower number of PNIPMAM-water hydrogen bonds as compared to ...
  18. [18]
    Using Implicit-Solvent Potentials to Extract Water Contributions to ...
    Jul 25, 2023 · For association in lower-critical-solution-temperature (LCST) systems, water reorganization entropy dominates the favorable free-energy change, ...
  19. [19]
    Hydrophobic hydration of poly-N-isopropyl acrylamide - Nature
    Mar 14, 2014 · Several investigations have shown that the LCST of PNiPAM in water is almost independent of PNiPAM molecular weight and concentration, yet ...Missing: penalty | Show results with:penalty
  20. [20]
    Effect of Hydrophobic Interactions on Lower Critical Solution ... - MDPI
    Jun 4, 2019 · While the hydrophobic DMA lowers the LCST especially the onset, polymerization degree has an important but smaller influence over all the whole LCST range.
  21. [21]
    Effects of stereochemistry and copolymerization on the LCST of ...
    Jan 20, 2017 · ... polymer conformational entropy upon collapse. This is the well established entropy driven mechanism of LCST polymer collapse.20 Additionally ...
  22. [22]
    Entropy-driven polymer collapse on increasing the temperature
    Apr 15, 2024 · Poly(N-vinylcaprolatam) chains undergo in water a temperature-induced, reversible and endothermic collapse. · The process is entropy-driven, and ...Missing: flexibility | Show results with:flexibility
  23. [23]
    Study of the Polymer Length Dependence of the Single Chain ...
    Aug 10, 2012 · We have performed simulations to examine the chain length dependence of the LCST transition and to characterize the single chain structure.
  24. [24]
    Effects of Salt on the Lower Critical Solution Temperature of Poly (N ...
    Aug 7, 2025 · The presence of salt in the solution will shift its LCST, typically to a lower temperature. This LCST shift follows the so-called Hofmeister ...
  25. [25]
    Cosolvent Control of Lower and Upper Critical Solution Behavior in ...
    Jul 1, 2025 · We report that polar cosolvent–water mixtures offer a unique approach to controlling the liquid–liquid phase separation (LLPS) of polyelectrolyte complex ...
  26. [26]
    A cosolvent surfactant mechanism affects polymer collapse in ... - NIH
    Nov 11, 2020 · Amphiphilic cosolvents can cause both swelling and collapse of stimuli responsive polymers, but the mechanisms governing this behavior remain ...
  27. [27]
    Polymers Exhibiting Lower Critical Solution Temperatures as a ...
    Nov 20, 2020 · Heating any form of thermoreversible gel past a critical temperature results in sol–gel transition, caused by an overall increase in hydrophobic ...<|control11|><|separator|>
  28. [28]
    Thermoresponsive polymers with LCST transition - RSC Publishing
    Aug 29, 2023 · The tunability of LCST makes the thermoresponsive polymers ideal for use in physiological temperatures, typically 35–40 °C range. Fig.
  29. [29]
    Improving the Colloidal Stability of Temperature-Sensitive Poly(N ...
    Sep 25, 2018 · Poly(N-isopropylacrylamide) (PNIPAM) is a well-known temperature-responsive polymer with LCST at approximately 32 °C, (19) close to human body ...
  30. [30]
    Temperature‐responsive polymers with LCST in the physiological ...
    Apr 13, 2007 · Temperature-responsive polymers with LCST in the physiological range can also be synthesized from polymers with a LCST much higher or lower ...
  31. [31]
    Measurement of cloud point temperature in polymer solutions
    Jul 30, 2013 · To derive the cloud point curve of ternary polymer solutions, a turbidity measurement apparatus was designed, instrumented, and set up.INTRODUCTION · II. EXPERIMENTAL... · Sample vial · “QUASI-EQUILIBRIUM...
  32. [32]
    Phase separation of triethylamine and water in native and ...
    Sep 21, 2017 · First, Ton decreases from 17.4 °C in the bulk to 3.9 °C in 4 nm native silica pores. The second trend is observed in the decrease of enthalpy.INTRODUCTION · Materials · Sample preparation · Native silica pores
  33. [33]
    System triethylamine–water: the equilibrium diagram and some ...
    Density and viscosity measurements at 16.0 °C, slightly below the critical solution temperature of 18.3 °C, show marked non-ideality. The negative volume of ...Missing: primary | Show results with:primary
  34. [34]
    Thermodynamic properties of triethylamine + water liquid mixture ...
    Apr 15, 2009 · Below the LCST, the hydrogen bonding between different molecules makes the mixture miscible. With increasing temperature, the hydrogen bonding ...
  35. [35]
    [PDF] Dynamics and Microstructures of Nicotine/Water Binary Mixtures ...
    Sep 19, 2018 · The LCST for the nicotine/water binary system is ∼60 °C. For the three temperatures at or above LCST, the ratio is the same within error.
  36. [36]
    Full article: Understanding the Fundamentals of Perfluorocarbons ...
    Jul 11, 2009 · 40°C lower critical solution temperature in hexane ( Table 2 ). Table 2 also illustrates the fact that all PFCs are not equivalent.
  37. [37]
    [PDF] Perfluorocarbon Emulsion Contrast Agents: A Mini Review - Frontiers
    Jan 10, 2022 · Although perfluorocarbons have low hydrocarbon affinity, halogen ... Critical solution temperature (n-hexane, °C). -20.0. Solubility in ...
  38. [38]
    Lower critical solution temperature (LCST) phase behaviour of an ...
    May 26, 2016 · LCST phase behaviour of an imidazolium-based ionic liquid is reported, which can be controlled by concentration, the choice of cation, anion and solvent.
  39. [39]
    Unusual LCST-type behaviour found in binary mixtures of choline ...
    Most of the systems exhibit unusual liquid–liquid demixing behavior characterized by the existence of lower critical solution temperatures (LCSTs). To the best ...
  40. [40]
    Smart ionic liquid/water mixture system with dual stimuli-response to ...
    Jul 11, 2022 · We synthesized five different polyetheramine-fatty acids (PEFA) ionic liquids (ILs), possessing the dual stimuli-responsive ability to temperature and CO 2.
  41. [41]
    Performance enhancement of aqueous ionic liquids with lower ...
    Nov 20, 2024 · LCST ionic liquids are typically comprised of a cation and anion with hydrophilic and hydrophobic functional groups that lead to a temperature- ...
  42. [42]
    Combinatorial discovery of small-molecule 1,2,3-triazolium ionic ...
    Oct 26, 2020 · LCST and upper critical solution temperature (UCST) systems are two typical phase behaviors of TILs, in which LCST is far less common than UCST.
  43. [43]
  44. [44]
    Concentration and Temperature Dependence of the Interaction ...
    Sep 21, 2022 · In contrast to the classical Flory–Huggins theory, χ values determined from RPA analysis were shown to increase with polymer volume fraction, ...
  45. [45]
  46. [46]
  47. [47]
    Lower critical solution temperature behavior of poly(2-chloroethyl ...
    Sep 8, 2010 · In some organic solvents, the polymer presents LCST-type thermoresponsive phase behavior under mild conditions. The clouding temperature of ...
  48. [48]
    Pressure Dependence of Upper and Lower Critical Solution ...
    Pressure Effects on Polymer Coil−Globule Transitions near an LCST. ... Lower Critical (LCST) and Upper Critical (UCST) Solution Temperatures. 2009 ...<|control11|><|separator|>
  49. [49]
    Characterization of the LCST behaviour of aqueous poly(N ...
    Polymer communication. Characterization of the LCST behaviour of aqueous poly(N-isopropylacrylamide) solutions by thermal and cloud point techniques.
  50. [50]
    When a Small Amount of Comonomer Is Enough - MDPI
    The TCP and TCL of all the copolymers exhibit a linear dependence on the mPEGMA1100 content within the investigated composition range. Thus, a linear ...Missing: T_LCST | Show results with:T_LCST
  51. [51]
    Macromolecules - ACS Publications - American Chemical Society
    Group contribution methods for predicting polymer-polymer miscibility from heats of mixing of liquids. 1. Comparison of the modified Guggenheim quasi ...
  52. [52]
    Effects of end group polarity and molecular weight on the lower ...
    Jan 11, 2006 · In the case of PNIPAM, reports variously say that LCSTs increase, decrease, or remain unchanged with molecular weight. Works by Fujishige et al.
  53. [53]
  54. [54]
    Influence of Additives on the Thermoresponsive Behavior of ...
    Aug 6, 2025 · For example, the addition of salt is known to reduce the LCST of PNIPAM microgels by disrupting the hydration structure surrounding the PNIPAM ...
  55. [55]
    A QSPR study for predicting θ(LCST) and θ(UCST) in binary ...
    Mar 5, 2023 · In this work, quantitative structure property relationship (QSPR) modeling was employed to predict θ(LCST) and θ(UCST) (LCST and UCST at the limit of infinite ...Missing: post- | Show results with:post-
  56. [56]
    All-Atom Molecular Dynamics Simulations of the Temperature ...
    PGE, which is actively studied in biomedicine, requires strict temperature control. If the LCST temperature can be predicted by MD simulations, the polymer ...
  57. [57]
    Prediction of the x Parameter by the Solubility ... - ACS Publications
    Lower critical solution temperature (LCST) and theta temperature of aqueous solutions of nonionic surface active agents of various polyoxyethylene chain lengths ...
  58. [58]
    Prediction of the lower critical solution temperature of poly(N ...
    The LCST of all the polymers decreased with increasing salt concentration. This is the first study to predict the LCST using support vector regression.Missing: history early<|control11|><|separator|>
  59. [59]
    Fingerprint applicable for machine learning tested on LCST ...
    Sep 20, 2023 · We introduce the polymer fingerprint, a way of representing polymers from common experimental characterization data such as size-exclusion chromatography.
  60. [60]
    Machine Learning‐Aided Prediction and Construction of a ...
    Sep 10, 2023 · This study demonstrates the usefulness of data-driven ML techniques for polymer research with the rapid and accurate prediction of polymer properties.
  61. [61]
    Poly(N-isopropylacrylamide)-Based Hydrogels for Biomedical ...
    The hydrogen bond between water molecules and PNIPAM is weaker due to the temperature rising above LCST, leading to the formation of an unstable solution.Poly(n-Isopropylacrylamide)-Based... · 7. Pnipam-Based Hydrogels In... · 9. Pnipams Hydrogel In...<|control11|><|separator|>
  62. [62]
    Design of Temperature-Responsive Cell Culture Surfaces for Cell ...
    This review introduces concepts of cell sheet engineering, followed by designs for the fabrication of various types of temperature-responsive cell culture ...
  63. [63]
    The construction of elastin-like polypeptides and their applications in ...
    Nov 11, 2023 · Elastin-like polypeptides (ELPs) are thermally responsive biopolymers derived from natural elastin. These peptides have a low critical ...
  64. [64]
    Engineering Temperature-Responsive Polymer Nanoparticles that ...
    Dec 27, 2023 · Poly(N-isopropylacrylamide) (pNIPAm) undergoes a hydrophilicity/hydrophobicity change around its lower critical solution temperature (LCST).
  65. [65]
    Monomer/solvent separation and recycle process for propylene ...
    The two phases are separated in the downstream Lower Critical Solution Temperature separator, also referred to as a “LCST” or a “gravimetric separator” or ...
  66. [66]
    Thermoresponsive polymers for environmentally friendly ...
    Figure: Thermoresponsive polymers with LCST-behavior are soluble in water below the LCST, enabling green electrospinning under controlled environmental ...
  67. [67]
    Polymerization of a new thermo-responsive copolymer with N ...
    Oct 12, 2018 · A new TR polymer PVBAm with suitable LCST is synthesized and can be recycled with high recovery, and a new recyclable ATPS is designed using a ...
  68. [68]
    Current applications of different type of aqueous two-phase systems
    Dec 28, 2015 · In recent year, aqueous two-phase system (ATPS) has become a proven tool used in separation and purification technology.
  69. [69]
    Thermodynamic properties of lower critical solution temperature ...
    LCST mixtures are homogeneous below the LCST, separating above it. They need more negative partial molar enthalpy and entropy for improved performance.Missing: definition | Show results with:definition
  70. [70]
    Expeditious Discovery of Small-Molecule Thermoresponsive Ionic ...
    Sep 27, 2023 · Therefore, we can sum up the causes of the ionic liquids of UCST and LCST. ... CO2 capture by a task-specific ionic liquid. J. Am. Chem ...
  71. [71]
  72. [72]
    A green solvent-to-polymer upgrading approach to water-soluble ...
    Oct 7, 2022 · In this study, the green solvents EL and DML were employed as platform chemicals to develop novel sustainable LCST-type acrylic (co)polymers ...
  73. [73]
    Sustainable 3D printing by reversible salting-out effects with ... - Nature
    May 9, 2024 · A facile and sustainable 3D printing technique is developed by utilizing a reversible salting-out effect.