Nonequilibrium relaxation of soft responsive colloids

José López-Molina Department of Applied Physics, University de Granada, Campus Fuentenueva S/N, 18071 Granada, Spain    Sebastien Groh Physikalisches Institut, Albert-Ludwigs-Universität Freiburg, Hermann-Herder Straße 3, D-79104 Freiburg, Germany    Joachim Dzubiella [email protected] Physikalisches Institut, Albert-Ludwigs-Universität Freiburg, Hermann-Herder Straße 3, D-79104 Freiburg, Germany Cluster of Excellence livMatS @ FIT - Freiburg Center for Interactive Materials and Bioinspired Technologies, Albert-Ludwigs-Universität Freiburg, D-79110 Freiburg, Germany    Arturo Moncho-Jordá [email protected] Department of Applied Physics, University de Granada, Campus Fuentenueva S/N, 18071 Granada, Spain Institute Carlos I for Theoretical and Computational Physics, University de Granada, Campus Fuentenueva S/N, 18071 Granada, Spain
(July 1, 2024)
Abstract

Stimuli-responsive macromolecules display large conformational changes during their dynamics, sometimes switching states, and are an integral property for the development of soft functional materials. Here, we introduce a mean-field dynamical density functional theory (DDFT) for a model of responsive colloids (RCs) to study the nonequilibrium dynamics of a colloidal dispersion in time-dependent external fields, with a focus on the coupling of translational and conformational dynamics during their relaxation. Specifically, we consider soft Gaussian particles with a bimodal size distribution between two confining walls with time-dependent (switching-on and off) external gravitational and osmotic fields. We find a rich relaxation behavior of the systems in excellent agreement with particle-based Brownian dynamics (BD) computer simulations. In particular, we find time-asymmetric relaxations of integrated observables (wall pressures, mean size, liquid center-of-mass) for activation/deactivation of external potentials, respectively, which are tuneable by the ratio of translational and conformational diffusion time scales. Our work paves thus the way for studying the nonequilibrium relaxation dynamics of complex soft matter with multiple degrees of freedom and hierarchical relaxations.

preprint: AIP/123-QED

I Introduction

Responsive systems have garnered considerable interest in the realm of soft matter science due to their dynamic nature and the ability to adapt to external stimuli. Stuart et al. (2010) These systems, comprising responsive colloids and macromolecules, exhibit remarkable adaptability, controlling their properties—such as the internal conformation of the particle, Onuchic, Luthey-Schulten, and Wolynes (1997); Wu and Wang (1998); Cho, Levy, and Wolynes (2006); Choi et al. (2011a) size, Denton and Schmidt (2002); Urich and Denton (2016); Brijitta and Schurtenberger (2019); Baul and Dzubiella (2021); Baul et al. (2021a); Lin, Rotenberg, and Dzubiella (2020) shape, Meng and Li (2013); Lee et al. (2019); Brijitta and Schurtenberger (2019); Lim and Denton (2014, 2016a, 2016b); Harrer et al. (2019); Del Monte and Zaccarelli (2024); Elancheliyan et al. (2022) charge density, Brijitta and Schurtenberger (2019); Weyer and Denton (2018) electric dipole Cao and Berne (1993) and orientation, Calef and Wolynes (1983); Chandra and Bagchi (1990, 1988) among others - in response to environmental changes. Such responsiveness, originating from their internal degrees of freedom (DoFs), allows for a nuanced interaction with surrounding particles and external fields, leading to significant alterations in their internal and collective dynamical properties,Motornov et al. (2007, 2011); Kalaitzidou and Crosby (2008); Stuart et al. (2010); Cao and Berne (1993); Cheung, Klimov, and Thirumalai (2005); Zhou, Rivas, and Minton (2008); Hong and Gierasch (2010); Dupuis, Holmstrom, and Nesbitt (2014); Shin, Cherstvy, and Metzler (2015); Baul et al. (2021b); Huang et al. (2016) even leading to multi-relaxation time scales. Garbin et al. (2015)

An important system of RCs is the one for which the particle size (that could represent the radius of gyration of a linear polymer coil, Denton and Schmidt (2002); Vettorel, Besold, and Kremer (2010) or the diameter of a microgel particle Winkler, Fedosov, and Gompper (2014); Moncho-Jordá and Dzubiella (2016); Karg et al. (2019); Brijitta and Schurtenberger (2019); Rovigatti et al. (2019); Scotti et al. (2019a, b); Scotti (2021)) is the internal property that couples to the center of mass translational degrees of freedom: particles not only move, but also are able to swell/shrink in response to external stimuli such as changes in the solvent pH, salt concentration and temperature.Murray and Snowden (1995); Saunders and Vincent (1999); Fernandez-Nieves et al. (2011); Zhou, Tang, and Wu (2015); Bochenek et al. (2022) In addition, particle concentration can also provoke the squeezing of the particles, leading to interpenetration, deformation and compression. Nikolov, Fernandez-Nieves, and Alexeev (2020); Gnan and Zaccarelli (2019) Precise control over the interplay between size localization and dynamics is paramount for achieving targeted functionality at specific locations and rates. Notable examples are the local modulation of uptake and release kinetics in soft polymer-based nanocarriers, such as microgels for local control of catalysis Roa et al. (2017, 2018); Kanduč et al. (2020) or drug release. Hamidi, Azadi, and Rafiei (2008); Saunders et al. (2009); Vinogradov (2010); Moncho-Jordá et al. (2019, 2020)

Our understanding of soft colloids has been significantly enhanced through developments in equilibrium Density Functional Theory (DFT), which has provided a solid framework for studying the structure and phase behavior of soft materials under external potentials. Classical DFT has been successfully extended to explicitly include varying particle sizes as dynamic variables, describing how microscopic interactions and external potentials affect the equilibrium properties of these systems and enabling the exploration of size-dependent phenomena within polydisperse systems. Pagonabarraga, Cates, and Ackland (2000); Buzzacchi, Pagonabarraga, and Wilding (2004); Denton and Schmidt (2002); Pagonabarraga and Cates (2001) Recently, we demonstrated explicitly for a model of responsive colloids (RC)s with size polydispersity Urich and Denton (2016); Baul and Dzubiella (2021); Baul et al. (2021a); Lin, Rotenberg, and Dzubiella (2020) how to employ DFT functionals (in RC-DFT) to study and control the localization of the (size) property in space by external fields. Moncho-Jordá, Groh, and Dzubiella (2024); Moncho-Jordá, Göth, and Dzubiella (2023)

However, many interesting behaviors of responsive systems occur out of equilibrium. The dynamical nature of these systems is not only a testament to their adaptability but also to their potential in real applications. In this context, dynamical density functional theory (DDFT) serves as a powerful tool for exploring non-equilibrium processes by modeling the time evolution of the particle density distribution, ρ(𝐫,t)𝜌𝐫𝑡\rho(\mathbf{r},t)italic_ρ ( bold_r , italic_t ), driven by diffusive, overdamped Brownian dynamics in the presence of external fields and particle interactions. te Vrugt, Löwen, and Wittkowski (2020); Marconi and Tarazona (1999); Archer and Evans (2004a); Royall et al. (2007); te Vrugt, Löwen, and Wittkowski (2020) The theory can also be modified to incorporate reactions or switching te Vrugt, Bickmann, and Wittkowski (2020); Moncho-Jordá and Dzubiella (2020); Bley, Dzubiella, and Moncho-Jordá (2021); Bley et al. (2022) Crucially, it leverages the adiabatic approximation, assuming that the correlations in a non-equilibrium state are akin to those in equilibrium. te Vrugt, Löwen, and Wittkowski (2020) This theory adapts the equilibrium concepts of DFT to non-equilibrium scenarios, predicting how responsive systems evolve over time. In contrast to conventional polydisperse systems, Pagonabarraga, Cates, and Ackland (2000); Buzzacchi, Pagonabarraga, and Wilding (2004) internal degrees of freedom are also able to change during the nonequilibrium process. A systematic study on the coupled dynamics of translation and internal dynamics in the relaxation of a colloidal system as well as the appropriate DDFT is still missing in literature.

In this work, we present an extension of classical DDFT to responsive systems (denoted by RC-DDFT) that allows us to efficiently investigate the dynamical relaxation in the presence of coupled DoFs under applied external potentials. In particular, we consider soft responsive colloids (RC), for which the size of the particles (σ𝜎\sigmaitalic_σ) changes in response to the interactions with the rest of particles or with an applied external potential. This external potential depends on the position and also the size of the particle, i.e. uext(𝐫,σ)subscript𝑢ext𝐫𝜎u_{\textmd{ext}}(\mathbf{r},\sigma)italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , italic_σ ), making the colloidal system inhomogeneous in terms of position and size. In addition to the external potential, this system of responsive colloids also requires the knowledge of the free energy landscape for the particle size, ψ(σ)=kTlnp(σ)𝜓𝜎𝑘𝑇𝑝𝜎\psi(\sigma)=-kT\ln p(\sigma)italic_ψ ( italic_σ ) = - italic_k italic_T roman_ln italic_p ( italic_σ ) (where p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ) represents the parent size-distribution of a single RC), which acts as an additional external potential controlling the size fluctuations. Lin, Rotenberg, and Dzubiella (2020) Here, we focus on a system formed by bistable particles for which p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ) is described by a generic Landau-like bimodal size distribution, Sato Matsuo and Tanaka (1988); Suzuki and Suzuki (1995) so particle size fluctuate between two states (big and small) separated by an energy barrier. This particular two-states behavior is relevant in the conformation of many biological or functional macromolecules, such as folded/unfolded or globule/coil transitions of proteins and polymers. Plaxco et al. (2000); Murnen et al. (2012); Zeng, Ruff, and Pappu (2022); Choi et al. (2011b); Dhiman, Jain, and George (2017); Dupuis, Holmstrom, and Nesbitt (2014); Kang et al. (2015) For this system, we analyse the time evolution of the one-body density profile, ρ(𝐫,σ;t)𝜌𝐫𝜎𝑡\rho(\mathbf{r},\sigma;t)italic_ρ ( bold_r , italic_σ ; italic_t ), after sudden activation/deactivation of the external field, and explore the non-equilibrium transient dynamics resulting of the interplay between structural relaxation and size relaxation.

This paper is organized as follows. In Sec. II we describe the main statistical mechanics equations, discuss DFT and generalize its dynamical counterpart to deal with non-equilibrium systems of responsive colloids under external potentials (RC-DDFT). We also introduce in this Section a mean field model for soft Gaussian colloids with a bimodal distribution of states. Brownian dynamic simulations of RCs are explained in Sec. III. Sec. IV presents the results and discussion of the non-equilibrium dynamics of RCs under the activation/deactivation non-equilibrium processes for two representative external potentials: gravitation and osmotic. We specially focus on investigating the appearance of non-equilibrium transient dynamics states that arise due to the interplay between translational diffusion and particle swelling/shrinking. Finally, in Sec. V we present the main conclusions of our work.

II Theoretical background

II.1 Theoretical modelling of responsive colloids (RCs)

In the following, we briefly recall the most basic statistical relations between distributions and averages for the RC model. Lin, Rotenberg, and Dzubiella (2020); Moncho-Jordá, Göth, and Dzubiella (2023) As common in the theory of liquids, we assume a (isotropic) distance-dependent pair potential for the RC liquid. For RCs, however, the particle-particle pair potential not only depends on the positions 𝐫isubscript𝐫𝑖\mathbf{r}_{i}bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and 𝐫jsubscript𝐫𝑗\mathbf{r}_{j}bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT of both particles i𝑖iitalic_i and j𝑗jitalic_j, but also has explicit dependence on the size of both interacting particles, σisubscript𝜎𝑖\sigma_{i}italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and σjsubscript𝜎𝑗\sigma_{j}italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT, that is u(|𝐫i𝐫j|;σi,σj)𝑢subscript𝐫𝑖subscript𝐫𝑗subscript𝜎𝑖subscript𝜎𝑗u(|\mathbf{r}_{i}-\mathbf{r}_{j}|;\sigma_{i},\sigma_{j})italic_u ( | bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | ; italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ). The external potential can be expressed as uext(𝐫i,σi)subscript𝑢extsubscript𝐫𝑖subscript𝜎𝑖u_{\textmd{ext}}(\mathbf{r}_{i},\sigma_{i})italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ). Hence, the total potential energy of N𝑁Nitalic_N responsive particles, can be expressed as

U=iψ(σi)+12ijiu(|𝐫i𝐫j|,σi,σj)+iuext(𝐫i,σi),𝑈subscript𝑖𝜓subscript𝜎𝑖12subscript𝑖subscript𝑗𝑖𝑢subscript𝐫𝑖subscript𝐫𝑗subscript𝜎𝑖subscript𝜎𝑗subscript𝑖subscript𝑢extsubscript𝐫𝑖subscript𝜎𝑖U=\sum_{i}\psi(\sigma_{i})+\frac{1}{2}\sum_{i}\sum_{j\neq i}u(|\mathbf{r}_{i}-% \mathbf{r}_{j}|,\sigma_{i},\sigma_{j})+\sum_{i}u_{\textmd{ext}}(\mathbf{r}_{i}% ,\sigma_{i}),italic_U = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_ψ ( italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j ≠ italic_i end_POSTSUBSCRIPT italic_u ( | bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (1)

where ψ𝜓\psiitalic_ψ is the energy landscape for the size σ𝜎\sigmaitalic_σ of an isolated particle. The inhomogeneous properties of an RC fluid immersed in a external potential are fully described by the particle density distribution, ρ(𝐫,σ)𝜌𝐫𝜎\rho(\mathbf{r},\sigma)italic_ρ ( bold_r , italic_σ )Lin, Rotenberg, and Dzubiella (2020); Moncho-Jordá, Göth, and Dzubiella (2023) defined in such a way that integration over the 4444 coordinates (3333 translational ones and the size) provides the total number of particles in the system,

V𝑑𝐫𝑑σρ(𝐫,σ)=N,subscript𝑉differential-d𝐫differential-d𝜎𝜌𝐫𝜎𝑁\int_{V}d\mathbf{r}\int d\sigma\rho(\mathbf{r},\sigma)=N,∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_d bold_r ∫ italic_d italic_σ italic_ρ ( bold_r , italic_σ ) = italic_N , (2)

so ρ(𝐫,σ)𝜌𝐫𝜎\rho(\mathbf{r},\sigma)italic_ρ ( bold_r , italic_σ ) is measured in units of length-4.

In the limit of very low particle concentrations and negligible external potential, the one-body density converges to limρ00limuext0ρ(𝐫,σ)=ρ0p(σ)subscriptsubscript𝜌00subscriptsubscript𝑢ext0𝜌𝐫𝜎subscript𝜌0𝑝𝜎\lim_{\rho_{0}\rightarrow 0}\ \lim_{u_{\textmd{ext}}\rightarrow 0}\rho(\mathbf% {r},\sigma)=\rho_{0}p(\sigma)roman_lim start_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT roman_lim start_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT italic_ρ ( bold_r , italic_σ ) = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_p ( italic_σ ), where ρ0=N/Vsubscript𝜌0𝑁𝑉\rho_{0}=N/Vitalic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_N / italic_V is the bulk particle density, and p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ) is the so called parent distribution, defined as the size distribution of a single isolated responsive particle that does not interact with external forces or with another particles, which is normalized to unity, p(σ)𝑑σ=1𝑝𝜎differential-d𝜎1\int p(\sigma)d\sigma=1∫ italic_p ( italic_σ ) italic_d italic_σ = 1. We can express p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ) in terms of the free energy landscape as

p(σ)=p0eβψ(σ),𝑝𝜎subscript𝑝0superscript𝑒𝛽𝜓𝜎p(\sigma)=p_{0}e^{-\beta\psi(\sigma)},italic_p ( italic_σ ) = italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_β italic_ψ ( italic_σ ) end_POSTSUPERSCRIPT , (3)

where β=1/(kBT)𝛽1subscript𝑘𝐵𝑇\beta=1/(k_{B}T)italic_β = 1 / ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) (T𝑇Titalic_T is the absolute temperature and kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT the Boltzmann constant), and p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is a constant prefactor to fulfill the normalization of p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ). For higher particle concentrations or in the presence of external fields, the single-particle property distribution will change: We denote by f(σ)=N(σ)/N𝑓𝜎𝑁𝜎𝑁f(\sigma)=N(\sigma)/Nitalic_f ( italic_σ ) = italic_N ( italic_σ ) / italic_N the emergent probability distribution, where N(σ)dσ𝑁𝜎𝑑𝜎N(\sigma)d\sigmaitalic_N ( italic_σ ) italic_d italic_σ is the number of particles with internal property within [σ,σ+dσ]𝜎𝜎𝑑𝜎\left[\sigma,\sigma+d\sigma\right][ italic_σ , italic_σ + italic_d italic_σ ] in the system, and N𝑁Nitalic_N the total number of particles. In particular, in the absence of any external field, the one-body particle density may be expressed as limuext0ρ(𝐫,σ)=ρ0f(σ)subscriptsubscript𝑢ext0𝜌𝐫𝜎subscript𝜌0𝑓𝜎\lim_{u_{\textmd{ext}}\rightarrow 0}\rho(\mathbf{r},\sigma)=\rho_{0}f(\sigma)roman_lim start_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT → 0 end_POSTSUBSCRIPT italic_ρ ( bold_r , italic_σ ) = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_f ( italic_σ ).

II.2 Dynamical density functional theory for RCs

II.2.1 Equilibrium DFT prerequisites

Let us recall first the treatment of the RC model in the framework of equilibrium DFT. The inhomogeneous free energy functional of a RC fluid immersed in the external potential uext(r,σ)subscript𝑢extr𝜎u_{\textmd{ext}}(\mathrm{r},\sigma)italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( roman_r , italic_σ ) is given by Pagonabarraga, Cates, and Ackland (2000); Lin, Rotenberg, and Dzubiella (2020); Moncho-Jordá, Göth, and Dzubiella (2023)

F[ρ(𝐫,σ)]𝐹delimited-[]𝜌𝐫𝜎\displaystyle F[\rho(\mathbf{r},\sigma)]italic_F [ italic_ρ ( bold_r , italic_σ ) ] =\displaystyle== kBT𝑑𝐫𝑑σρ(𝐫,σ)[ln(ρ(𝐫,σ)Λ3/p0)1]subscript𝑘𝐵𝑇differential-d𝐫differential-d𝜎𝜌𝐫𝜎delimited-[]𝜌𝐫𝜎superscriptΛ3subscript𝑝01\displaystyle k_{B}T\int d\mathbf{r}\int d\sigma\rho(\mathbf{r},\sigma)\big{[}% \ln(\rho(\mathbf{r},\sigma)\Lambda^{3}/p_{0})-1\big{]}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ∫ italic_d bold_r ∫ italic_d italic_σ italic_ρ ( bold_r , italic_σ ) [ roman_ln ( italic_ρ ( bold_r , italic_σ ) roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) - 1 ]
+\displaystyle++ 𝑑𝐫𝑑σρ(𝐫,σ)[uext(𝐫,σ)+ψ(σ)]differential-d𝐫differential-d𝜎𝜌𝐫𝜎delimited-[]subscript𝑢ext𝐫𝜎𝜓𝜎\displaystyle\int d\mathbf{r}\int d\sigma\rho(\mathbf{r},\sigma)[u_{\textmd{% ext}}(\mathbf{r},\sigma)+\psi(\sigma)]∫ italic_d bold_r ∫ italic_d italic_σ italic_ρ ( bold_r , italic_σ ) [ italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , italic_σ ) + italic_ψ ( italic_σ ) ]
+\displaystyle++ Fex[ρ(𝐫,σ)],subscript𝐹exdelimited-[]𝜌𝐫𝜎\displaystyle F_{\textmd{ex}}[\rho(\mathbf{r},\sigma)],italic_F start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT [ italic_ρ ( bold_r , italic_σ ) ] ,

where Λ=h/(2πmkBT)1/2Λsuperscript2𝜋𝑚subscript𝑘𝐵𝑇12\Lambda=h/(2\pi mk_{B}T)^{1/2}roman_Λ = italic_h / ( 2 italic_π italic_m italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT is the thermal wavelength. The first term of Eq. (II.2.1) is the ideal gas free-energy functional. The second term takes into account the interaction of the RC fluid with the external potential. Note that ψ(σ)𝜓𝜎\psi(\sigma)italic_ψ ( italic_σ ) also plays the role of an external potential for the particle size, i.e., it represents the energy cost implied in the swelling/shrinking of each responsive colloid. Finally, the third contribution is the excess free energy of the fluid that arises due to the existence of particle-particle interactions.

The grand canonical potential energy functional of a RC fluid is

Ω[ρ(𝐫,σ)]=F[ρ(𝐫,σ)]μ0𝑑𝐫𝑑σρ(𝐫,σ),Ωdelimited-[]𝜌𝐫𝜎𝐹delimited-[]𝜌𝐫𝜎subscript𝜇0differential-d𝐫differential-d𝜎𝜌𝐫𝜎\Omega[\rho(\mathbf{r},\sigma)]=F[\rho(\mathbf{r},\sigma)]-\mu_{0}\int d% \mathbf{r}\int d\sigma\rho(\mathbf{r},\sigma),roman_Ω [ italic_ρ ( bold_r , italic_σ ) ] = italic_F [ italic_ρ ( bold_r , italic_σ ) ] - italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ italic_d bold_r ∫ italic_d italic_σ italic_ρ ( bold_r , italic_σ ) , (5)

where μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the (constant) chemical potential of the RC fluid. The equilibrium density profile is the one that minimizes the grand canonical functional, δΩ/δρ(𝐫,σ)=0𝛿Ω𝛿𝜌𝐫𝜎0\delta\Omega/\delta\rho(\mathbf{r},\sigma)=0italic_δ roman_Ω / italic_δ italic_ρ ( bold_r , italic_σ ) = 0. Applying this functional differentiation to Eq. (5) with Eq. (II.2.1) and solving the resulting Euler-Lagrange equation for the particle density ρ(𝐫,σ)𝜌𝐫𝜎\rho(\mathbf{r},\sigma)italic_ρ ( bold_r , italic_σ ), we find

ρ(𝐫,σ)=qexp(βψ(σ)βuext(𝐫,σ)βμex(𝐫,σ)),𝜌𝐫𝜎𝑞𝛽𝜓𝜎𝛽subscript𝑢ext𝐫𝜎𝛽subscript𝜇ex𝐫𝜎\rho(\mathbf{r},\sigma)=q\exp\big{(}-\beta\psi(\sigma)-\beta u_{\textmd{ext}}(% \mathbf{r},\sigma)-\beta\mu_{\textmd{ex}}(\mathbf{r},\sigma)\big{)},italic_ρ ( bold_r , italic_σ ) = italic_q roman_exp ( - italic_β italic_ψ ( italic_σ ) - italic_β italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , italic_σ ) - italic_β italic_μ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r , italic_σ ) ) , (6)

where μex(𝐫,σ)subscript𝜇ex𝐫𝜎\mu_{\textmd{ex}}(\mathbf{r},\sigma)italic_μ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r , italic_σ ) is the excess chemical potential, given by the functional differentiation of the excess free energy, μex(𝐫,σ)=δFex/δρ(𝐫,σ)subscript𝜇ex𝐫𝜎𝛿subscript𝐹ex𝛿𝜌𝐫𝜎\mu_{\textmd{ex}}(\mathbf{r},\sigma)=\delta F_{\textmd{ex}}/\delta\rho(\mathbf% {r},\sigma)italic_μ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r , italic_σ ) = italic_δ italic_F start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT / italic_δ italic_ρ ( bold_r , italic_σ ), and q𝑞qitalic_q is a normalization constant that is obtained imposing conservation of the total number of particles, namely Eq. (2). Solutions of Eqs. (6) and (2) lead to the equilibrium position and size distribution ρeq(𝐫,σ)subscript𝜌eq𝐫𝜎\rho_{\textmd{eq}}(\mathbf{r},\sigma)italic_ρ start_POSTSUBSCRIPT eq end_POSTSUBSCRIPT ( bold_r , italic_σ ).

II.2.2 Dynamical DFT for a RC fluid

In non-equilibrium conditions, the one-body density distribution becomes also time-dependent, ρ(𝐫,σ;t)𝜌𝐫𝜎𝑡\rho(\mathbf{r},\sigma;t)italic_ρ ( bold_r , italic_σ ; italic_t ). For the case of responsive colloids, particles in the system do not only diffuse in the space, but can also can modify their their size in response to external interactions. To describe their dynamics we make the assumption that the property σ𝜎\sigmaitalic_σ of each particle also follows an overdamped diffusive dynamics. Following the prescription given in the work by Baul et al., Baul and Dzubiella (2021) we denote DTsubscript𝐷TD_{\textmd{T}}italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT as the translational diffusion coefficient, and Dσsubscript𝐷𝜎D_{\sigma}italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT the diffusion coefficient in the σ𝜎\sigmaitalic_σ-space. It is important to emphasize that DTsubscript𝐷TD_{\textmd{T}}italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT depends in general on the particle size (as for Stokes-Einstein), so DT(σ)subscript𝐷T𝜎D_{\textmd{T}}(\sigma)italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT ( italic_σ ) is a function of σ𝜎\sigmaitalic_σ.

The DDFT can be extended to RCs by defining a 4-dimensional vector 𝐱=(x,y,z,σ)(𝐫,σ)𝐱𝑥𝑦𝑧𝜎𝐫𝜎\mathbf{x}=(x,y,z,\sigma)\equiv(\mathbf{r},\sigma)bold_x = ( italic_x , italic_y , italic_z , italic_σ ) ≡ ( bold_r , italic_σ ), and a 4444-dimensional current 𝐉=(Jx,Jy,Jz,Jσ)(𝐉𝐫,Jσ)𝐉subscript𝐽𝑥subscript𝐽𝑦subscript𝐽𝑧subscript𝐽𝜎subscript𝐉𝐫subscript𝐽𝜎\mathbf{J}=(J_{x},J_{y},J_{z},J_{\sigma})\equiv(\mathbf{J}_{\mathbf{r}},J_{% \sigma})bold_J = ( italic_J start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ≡ ( bold_J start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT , italic_J start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) (note that this current has dimensions of length-3time-1). Analogously, we can write a 4444-component nabla operator 𝐱=(/x,/y,/z,/σ)(𝐫,σ)subscript𝐱𝑥𝑦𝑧𝜎subscript𝐫subscript𝜎\nabla_{\mathbf{x}}=(\partial/\partial x,\partial/\partial y,\partial/\partial z% ,\partial/\partial\sigma)\equiv(\nabla_{\mathbf{r}},\nabla_{\sigma})∇ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT = ( ∂ / ∂ italic_x , ∂ / ∂ italic_y , ∂ / ∂ italic_z , ∂ / ∂ italic_σ ) ≡ ( ∇ start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT , ∇ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ). The time evolution of the density profile is given by

ρ(𝐱,t)t=𝐱𝐉(𝐱,t)=r𝐉𝐫Jσσ𝜌𝐱𝑡𝑡subscript𝐱𝐉𝐱𝑡subscript𝑟subscript𝐉𝐫subscript𝐽𝜎𝜎\frac{\partial\rho(\mathbf{x},t)}{\partial t}=-\nabla_{\mathbf{x}}\cdot\mathbf% {J}(\mathbf{x},t)=-\nabla_{r}\cdot\mathbf{J}_{\mathbf{r}}-\frac{\partial J_{% \sigma}}{\partial\sigma}divide start_ARG ∂ italic_ρ ( bold_x , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG = - ∇ start_POSTSUBSCRIPT bold_x end_POSTSUBSCRIPT ⋅ bold_J ( bold_x , italic_t ) = - ∇ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ⋅ bold_J start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT - divide start_ARG ∂ italic_J start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_σ end_ARG (7)

The components of the current 𝐉𝐉\mathbf{J}bold_J are

{𝐉𝐫=DT(σ)ρ(𝐫,σ;t)𝐫[βμ(𝐫,σ;t)]Jσ=Dσρ(𝐫,σ;t)βμ(𝐫,σ;t)σcasessubscript𝐉𝐫subscript𝐷T𝜎𝜌𝐫𝜎𝑡subscript𝐫𝛽𝜇𝐫𝜎𝑡otherwisesubscript𝐽𝜎subscript𝐷𝜎𝜌𝐫𝜎𝑡𝛽𝜇𝐫𝜎𝑡𝜎otherwise\begin{cases}\mathbf{J}_{\mathbf{r}}=-D_{\textmd{T}}(\sigma)\rho(\mathbf{r},% \sigma;t)\nabla_{\mathbf{r}}[\beta\mu(\mathbf{r},\sigma;t)]\\ J_{\sigma}=-D_{\sigma}\rho(\mathbf{r},\sigma;t)\frac{\partial\beta\mu(\mathbf{% r},\sigma;t)}{\partial\sigma}\end{cases}{ start_ROW start_CELL bold_J start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT = - italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT ( italic_σ ) italic_ρ ( bold_r , italic_σ ; italic_t ) ∇ start_POSTSUBSCRIPT bold_r end_POSTSUBSCRIPT [ italic_β italic_μ ( bold_r , italic_σ ; italic_t ) ] end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_J start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = - italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT italic_ρ ( bold_r , italic_σ ; italic_t ) divide start_ARG ∂ italic_β italic_μ ( bold_r , italic_σ ; italic_t ) end_ARG start_ARG ∂ italic_σ end_ARG end_CELL start_CELL end_CELL end_ROW (8)

where μ(𝐫,σ;t)𝜇𝐫𝜎𝑡\mu(\mathbf{r},\sigma;t)italic_μ ( bold_r , italic_σ ; italic_t ) is the non-equilibrium chemical potential. In order to calculate it, we make use of the adiabatic approximation, and assume that μ(𝐫,σ;t)𝜇𝐫𝜎𝑡\mu(\mathbf{r},\sigma;t)italic_μ ( bold_r , italic_σ ; italic_t ) is given by the functional derivative of the equilibrium free energy functional, μ=δF/δρ𝜇𝛿𝐹𝛿𝜌\mu=\delta F/\delta\rhoitalic_μ = italic_δ italic_F / italic_δ italic_ρ. Using Eq. (II.2.1), it leads to Moncho-Jordá, Groh, and Dzubiella (2024)

μ(𝐫,σ;t)𝜇𝐫𝜎𝑡\displaystyle\mu(\mathbf{r},\sigma;t)italic_μ ( bold_r , italic_σ ; italic_t ) =\displaystyle== kBTln(ρ(𝐫,σ;t)Λ3/p0)+uext(𝐫,σ)+ψ(σ)subscript𝑘𝐵𝑇𝜌𝐫𝜎𝑡superscriptΛ3subscript𝑝0subscript𝑢ext𝐫𝜎𝜓𝜎\displaystyle k_{B}T\ln\big{(}\rho(\mathbf{r},\sigma;t)\Lambda^{3}/p_{0}\big{)% }+u_{\textmd{ext}}(\mathbf{r},\sigma)+\psi(\sigma)italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T roman_ln ( italic_ρ ( bold_r , italic_σ ; italic_t ) roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , italic_σ ) + italic_ψ ( italic_σ ) (9)
+\displaystyle++ μex(𝐫,σ;t),subscript𝜇ex𝐫𝜎𝑡\displaystyle\mu_{\textmd{ex}}(\mathbf{r},\sigma;t),italic_μ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r , italic_σ ; italic_t ) ,

where μex=δFex/δρsubscript𝜇ex𝛿subscript𝐹ex𝛿𝜌\mu_{\textmd{ex}}=\delta F_{\textmd{ex}}/\delta\rhoitalic_μ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT = italic_δ italic_F start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT / italic_δ italic_ρ. Eqs. (7)-(9) define the new generalized dynamical density functional theory designed for responsive colloids (RC-DDFT).

II.3 Mean-field responsive DDFT of soft Gaussian responsive colloids confined in planar slits

In this section, we specify our particular system of responsive colloids, and the corresponding excess free energy model for it to be used in the RC-DDFT framework. In this work we focus on systems composed by soft interpenetrable RCs. We consider the following size-dependent Gaussian-core pair potential for the particle-particle interaction

βu(|𝐫𝐫|,σ,σ)=ϵijexp(4|𝐫𝐫|2/(σ+σ)2),𝛽𝑢𝐫superscript𝐫𝜎superscript𝜎subscriptitalic-ϵ𝑖𝑗4superscript𝐫superscript𝐫2superscript𝜎superscript𝜎2\beta u(|\mathbf{r}-\mathbf{r}^{\prime}|,\sigma,\sigma^{\prime})=\epsilon_{ij}% \exp\Big{(}-4|\mathbf{r}-\mathbf{r}^{\prime}|^{2}/(\sigma+\sigma^{\prime})^{2}% \Big{)},italic_β italic_u ( | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | , italic_σ , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = italic_ϵ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT roman_exp ( - 4 | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_σ + italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (10)

where ϵ>0italic-ϵ0\epsilon>0italic_ϵ > 0 is the (repulsive) interaction strength (in kBTsubscript𝑘𝐵𝑇k_{B}Titalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T units). It represents an estimate of the particle hardness: for small values of ϵitalic-ϵ\epsilonitalic_ϵ colloids are able to interpenetrate each other. On the contrary, large values of ϵitalic-ϵ\epsilonitalic_ϵ correspond to harder colloids, for which the energy penalty of overlap** is very high. Here, (σ+σ)/2𝜎superscript𝜎2(\sigma+\sigma^{\prime})/2( italic_σ + italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) / 2 plays the role of the interaction range. In fact, in this interaction model σ𝜎\sigmaitalic_σ represents the effective radius of the each RC.

The Gaussian-core interaction model is a well established coarse-grained description of polymer solutions, Likos (2001) as it has been shown to accurately describe the interaction between two isolated polymers immersed in a good solvent, for polymer of identical Bolhuis et al. (2001) and different length Dautenhahn and Hall (1994), both in homogeneous and the inhomogeneous conditions. Louis, Bolhuis, and Hansen (2000) For different choices of the interaction parameters one can obtain either a mixture exhibiting bulk fluid–fluid (macro)phase separation Finken, Hansed, and Louis (2003); Archer and Evans (2001); Archer (2005a); Archer and Evans (2004b) similarly to polymer blends, or alternatively, a mixture exhibiting microphase separation. Archer, Likos, and Evans (2004)

We assume that ϵ=2italic-ϵ2\epsilon=2italic_ϵ = 2 (in kBTsubscript𝑘𝐵𝑇k_{B}Titalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T units) for the interparticle interaction strength. This value represents fairly well the soft repulsion existing between linear polymers. Louis, Bolhuis, and Hansen (2000); Archer and Evans (2001) The equilibrium properties of such soft Gaussian particles described by Eq. (10) are well represented by a weakly correlated mean-field fluid over a surprisingly wide density and temperature range. Louis, Bolhuis, and Hansen (2000) This justifies the use of the mean-field approximation for the excess free-energy of the interacting RC fluid, given by

Fex=12V𝑑𝐫𝑑𝐫𝑑σ𝑑σρ(𝐫,σ)ρ(𝐫,σ)u(|𝐫𝐫|,σ,σ)subscript𝐹ex12subscriptdouble-integral𝑉differential-d𝐫differential-dsuperscript𝐫double-integraldifferential-d𝜎differential-dsuperscript𝜎𝜌𝐫𝜎𝜌superscript𝐫superscript𝜎𝑢𝐫superscript𝐫𝜎superscript𝜎F_{\textmd{ex}}=\frac{1}{2}\iint_{V}d\mathbf{r}d\mathbf{r}^{\prime}\iint d% \sigma d\sigma^{\prime}\rho(\mathbf{r},\sigma)\rho(\mathbf{r}^{\prime},\sigma^% {\prime})u(|\mathbf{r}-\mathbf{r}^{\prime}|,\sigma,\sigma^{\prime})italic_F start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∬ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_d bold_r italic_d bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∬ italic_d italic_σ italic_d italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_ρ ( bold_r , italic_σ ) italic_ρ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_u ( | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | , italic_σ , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (11)

This approximation has been successfully used to reproduce the equilibrium and non-equilibrium properties of passive Archer (2005a, b) and active switching Gaussian colloids. Moncho-Jordá and Dzubiella (2020); Bley, Dzubiella, and Moncho-Jordá (2021, 2021); Bley et al. (2022)

Performing the functional differentiation and introducing it into Eq. (9) leads to the explicit expression for the non-equilibrium chemical potential of a mean-field RC fluid

μ(𝐫,σ;t)𝜇𝐫𝜎𝑡\displaystyle\mu(\mathbf{r},\sigma;t)italic_μ ( bold_r , italic_σ ; italic_t ) =\displaystyle== kBTln(ρ(𝐫,σ;t)Λ3/p0)+uext(𝐫,σ)+ψ(σ)subscript𝑘𝐵𝑇𝜌𝐫𝜎𝑡superscriptΛ3subscript𝑝0subscript𝑢ext𝐫𝜎𝜓𝜎\displaystyle k_{B}T\ln\big{(}\rho(\mathbf{r},\sigma;t)\Lambda^{3}/p_{0}\big{)% }+u_{\textmd{ext}}(\mathbf{r},\sigma)+\psi(\sigma)italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T roman_ln ( italic_ρ ( bold_r , italic_σ ; italic_t ) roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , italic_σ ) + italic_ψ ( italic_σ ) (12)
+\displaystyle++ 𝑑𝐫𝑑σρ(𝐫,σ;t)u(|𝐫𝐫|,σ,σ),differential-dsuperscript𝐫differential-dsuperscript𝜎𝜌superscript𝐫superscript𝜎𝑡𝑢𝐫superscript𝐫𝜎superscript𝜎\displaystyle\int d\mathbf{r}^{\prime}\int d\sigma^{\prime}\rho(\mathbf{r}^{% \prime},\sigma^{\prime};t)u(|\mathbf{r}-\mathbf{r}^{\prime}|,\sigma,\sigma^{% \prime}),∫ italic_d bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∫ italic_d italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_ρ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; italic_t ) italic_u ( | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | , italic_σ , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ,

which involves a convolution integral in the 𝐫𝐫\mathbf{r}bold_r coordinate.

In our work, we consider RC dispersions confined between two parallel hard walls separated by a distance L𝐿Litalic_L and subjected to one-dimensional external potentials, uext(z,σ)subscript𝑢ext𝑧𝜎u_{\textmd{ext}}(z,\sigma)italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( italic_z , italic_σ ), where z[0,L]𝑧0𝐿z\in[0,L]italic_z ∈ [ 0 , italic_L ] is the distance from the left wall. The rest of coordinates are assumed to run over the full range, x,y],[x,y\in]-\infty,\infty[italic_x , italic_y ∈ ] - ∞ , ∞ [ (infinite slit). In this case the density profiles are homogeneous in lateral directions and can be expressed as ρ(𝐫,σ;t)=ρ(z,σ;t)𝜌𝐫𝜎𝑡𝜌𝑧𝜎𝑡\rho(\mathbf{r},\sigma;t)=\rho(z,\sigma;t)italic_ρ ( bold_r , italic_σ ; italic_t ) = italic_ρ ( italic_z , italic_σ ; italic_t ), with the normalization

0L𝑑z𝑑σρ(z,σ;t)=N/S,superscriptsubscript0𝐿differential-d𝑧differential-d𝜎𝜌𝑧𝜎𝑡𝑁𝑆\int_{0}^{L}dz\int d\sigma\rho(z,\sigma;t)=N/S,∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_d italic_z ∫ italic_d italic_σ italic_ρ ( italic_z , italic_σ ; italic_t ) = italic_N / italic_S , (13)

where N/S𝑁𝑆N/Sitalic_N / italic_S is the prefixed number density per area S𝑆Sitalic_S, and the normalization is valid for every time t𝑡titalic_t.

Exploiting the planar symmetry to simplify the convolution integral involved in Eq. (12), we find the following equation for the non-equilibrium chemical potential Moncho-Jordá, Groh, and Dzubiella (2024)

μ(z,σ;t)𝜇𝑧𝜎𝑡\displaystyle\mu(z,\sigma;t)italic_μ ( italic_z , italic_σ ; italic_t ) =kBTln(ρ(z,σ;t)Λ3/p0)+uext(z,σ)+ψ(σ)absentsubscript𝑘𝐵𝑇𝜌𝑧𝜎𝑡superscriptΛ3subscript𝑝0subscript𝑢𝑒𝑥𝑡𝑧𝜎𝜓𝜎\displaystyle=k_{B}T\ln\big{(}\rho(z,\sigma;t)\Lambda^{3}/p_{0}\big{)}+u_{ext}% (z,\sigma)+\psi(\sigma)= italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T roman_ln ( italic_ρ ( italic_z , italic_σ ; italic_t ) roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) + italic_u start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT ( italic_z , italic_σ ) + italic_ψ ( italic_σ )
+\displaystyle++ πϵ4𝑑σ(σ+σ)20L𝑑zρ(z,σ;t)e4(zz)2(σ+σ)2.𝜋italic-ϵ4differential-dsuperscript𝜎superscript𝜎superscript𝜎2superscriptsubscript0𝐿differential-dsuperscript𝑧𝜌superscript𝑧superscript𝜎𝑡superscript𝑒4superscript𝑧superscript𝑧2superscript𝜎superscript𝜎2\displaystyle\frac{\pi\epsilon}{4}\int d\sigma^{\prime}(\sigma+\sigma^{\prime}% )^{2}\int_{0}^{L}dz^{\prime}\rho(z^{\prime},\sigma^{\prime};t)e^{-\frac{4(z-z^% {\prime})^{2}}{(\sigma+\sigma^{\prime})^{2}}}.divide start_ARG italic_π italic_ϵ end_ARG start_ARG 4 end_ARG ∫ italic_d italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_σ + italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_d italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_ρ ( italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; italic_t ) italic_e start_POSTSUPERSCRIPT - divide start_ARG 4 ( italic_z - italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_σ + italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_POSTSUPERSCRIPT . (14)

In addition to the interparticle interaction potential, we need to specify the parent distribution of the RC, p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ). In this work we consider bistable particles, for which the size fluctuates between two states of size σ1subscript𝜎1\sigma_{1}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and σ2subscript𝜎2\sigma_{2}italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, that can be referred as small and big. This implies considering a bimodal parent size distribution, with both peaks centered around these two states. To model this behavior, we choose a generic bimodal form using a symmetric double-Gaussian function

p(σ)=p022πτ2[exp((σσ1)22τ2)+exp((σσ2)22τ2)]𝑝𝜎subscript𝑝022𝜋superscript𝜏2delimited-[]superscript𝜎subscript𝜎122superscript𝜏2superscript𝜎subscript𝜎222superscript𝜏2p(\sigma)=\frac{p_{0}}{2\sqrt{2\pi\tau^{2}}}\Big{[}\exp\Big{(}-\frac{(\sigma-% \sigma_{1})^{2}}{2\tau^{2}}\Big{)}+\exp\Big{(}-\frac{(\sigma-\sigma_{2})^{2}}{% 2\tau^{2}}\Big{)}\Big{]}italic_p ( italic_σ ) = divide start_ARG italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 square-root start_ARG 2 italic_π italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG [ roman_exp ( - divide start_ARG ( italic_σ - italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) + roman_exp ( - divide start_ARG ( italic_σ - italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_τ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ] (15)

with σ1=0.63σ0subscript𝜎10.63subscript𝜎0\sigma_{1}=0.63\sigma_{0}italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.63 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, σ2=1.37σ0subscript𝜎21.37subscript𝜎0\sigma_{2}=1.37\sigma_{0}italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1.37 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represents a reference particle size that will be used as unit length for the rest of sizes and distances). In order to avoid nonphysical negative values and extremely large values of the particle size, the range of σ𝜎\sigmaitalic_σ has been limited to be σ[0.1σ0,2σ0]𝜎0.1subscript𝜎02subscript𝜎0\sigma\in[0.1\sigma_{0},2\sigma_{0}]italic_σ ∈ [ 0.1 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , 2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ].

The parameter τ𝜏\tauitalic_τ appearing in Eq. (15) provides the thickness of the size distribution around both peaks, and can be interpreted as the particle softness (conversely, τ1superscript𝜏1\tau^{-1}italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT represents the stiffness of the RC). In this work we use τ=0.2σ0𝜏0.2subscript𝜎0\tau=0.2\sigma_{0}italic_τ = 0.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. With this choice, the free energy barrier required to overcome to switch from one state to other is Δψ1kBTΔ𝜓1subscript𝑘𝐵𝑇\Delta\psi\approx 1k_{B}Troman_Δ italic_ψ ≈ 1 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T.

In our system, the σ𝜎\sigmaitalic_σ-dependence of the translational diffusion coefficient will be of the type of Stokes, DT(σ)=D0σ0/σsubscript𝐷T𝜎subscript𝐷0subscript𝜎0𝜎D_{\textmd{T}}(\sigma)=D_{0}\sigma_{0}/\sigmaitalic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT ( italic_σ ) = italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_σ, where D0subscript𝐷0D_{0}italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the diffusion coefficient of a particle of radius σ0subscript𝜎0\sigma_{0}italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In addition, a parameter α𝛼\alphaitalic_α is introduced to control the ratio between σ𝜎\sigmaitalic_σ-diffusion and translational diffusion, Dσ=αD0subscript𝐷𝜎𝛼subscript𝐷0D_{\sigma}=\alpha D_{0}italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT = italic_α italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We also define our diffusion time for either translation or size relaxation as τB=σ02/D0subscript𝜏Bsuperscriptsubscript𝜎02subscript𝐷0\tau_{\rm B}=\sigma_{0}^{2}/D_{0}italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

From ρ(z,σ;t)𝜌𝑧𝜎𝑡\rho(z,\sigma;t)italic_ρ ( italic_z , italic_σ ; italic_t ) and integrating over the σ𝜎\sigmaitalic_σ-coordinate we obtain the one-body number density distribution of the RC fluid in our quasi-1D system, namely

ρ(z;t)=𝑑σρ(z,σ;t).𝜌𝑧𝑡differential-d𝜎𝜌𝑧𝜎𝑡\rho(z;t)=\int d\sigma\rho(z,\sigma;t).italic_ρ ( italic_z ; italic_t ) = ∫ italic_d italic_σ italic_ρ ( italic_z , italic_σ ; italic_t ) . (16)

We analyze in our work a few integrated properties and monitor their time evolution: The center of mass location of the RC fluid is given by

z(t)=SN0Lz𝑑z𝑑σρ(z,σ;t).delimited-⟨⟩𝑧𝑡𝑆𝑁superscriptsubscript0𝐿𝑧differential-d𝑧differential-d𝜎𝜌𝑧𝜎𝑡\langle z(t)\rangle=\frac{S}{N}\int_{0}^{L}zdz\int d\sigma\rho(z,\sigma;t).⟨ italic_z ( italic_t ) ⟩ = divide start_ARG italic_S end_ARG start_ARG italic_N end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_z italic_d italic_z ∫ italic_d italic_σ italic_ρ ( italic_z , italic_σ ; italic_t ) . (17)

The mean size of the RC at position z𝑧zitalic_z is given by the normalized first moment of the distribution

σ(z;t)=1ρ(z;t)𝑑σρ(z,σ;t)σ.delimited-⟨⟩𝜎𝑧𝑡1𝜌𝑧𝑡differential-d𝜎𝜌𝑧𝜎𝑡𝜎\langle\sigma(z;t)\rangle=\frac{1}{\rho(z;t)}\int d\sigma\rho(z,\sigma;t)\sigma.⟨ italic_σ ( italic_z ; italic_t ) ⟩ = divide start_ARG 1 end_ARG start_ARG italic_ρ ( italic_z ; italic_t ) end_ARG ∫ italic_d italic_σ italic_ρ ( italic_z , italic_σ ; italic_t ) italic_σ . (18)

The (global) mean size of the RC is given by

σ(t)=SN0L𝑑z𝑑σρ(z,σ;t)σ.delimited-⟨⟩𝜎𝑡𝑆𝑁superscriptsubscript0𝐿differential-d𝑧differential-d𝜎𝜌𝑧𝜎𝑡𝜎\langle\sigma(t)\rangle=\frac{S}{N}\int_{0}^{L}dz\int d\sigma\rho(z,\sigma;t)\sigma.⟨ italic_σ ( italic_t ) ⟩ = divide start_ARG italic_S end_ARG start_ARG italic_N end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT italic_d italic_z ∫ italic_d italic_σ italic_ρ ( italic_z , italic_σ ; italic_t ) italic_σ . (19)

The pressure exerted on the left and right wall are given respectively by

Pleft(t)=ρ(0;t)kBTandPright(t)=ρ(L;t)kBT.formulae-sequencesubscript𝑃left𝑡𝜌0𝑡subscript𝑘𝐵𝑇andsubscript𝑃right𝑡𝜌𝐿𝑡subscript𝑘𝐵𝑇P_{\rm left}(t)=\rho(0;t)k_{B}T\ \ {\rm and}\ \ P_{\rm right}(t)=\rho(L;t)k_{B% }T.italic_P start_POSTSUBSCRIPT roman_left end_POSTSUBSCRIPT ( italic_t ) = italic_ρ ( 0 ; italic_t ) italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T roman_and italic_P start_POSTSUBSCRIPT roman_right end_POSTSUBSCRIPT ( italic_t ) = italic_ρ ( italic_L ; italic_t ) italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T . (20)

III Brownian Dynamics simulations of RCs

Brownian dynamics (BD) calculations are performed in the framework of the RC model.Lin, Rotenberg, and Dzubiella (2020) The discretized form of the BD equations for the translational degrees of freedom and for the internal property are given by:

{𝐫i(t+Δt)=𝐫i(t)+DTkBT𝐅T(i)(t)Δt+𝝃Tσi(t+Δt)=σi(t)+DσkBTFσ(i)(t)Δt+ξσcasessubscript𝐫𝑖𝑡Δ𝑡subscript𝐫𝑖𝑡subscript𝐷Tsubscript𝑘𝐵𝑇superscriptsubscript𝐅T𝑖𝑡Δ𝑡subscript𝝃Totherwisesubscript𝜎𝑖𝑡Δ𝑡subscript𝜎𝑖𝑡subscript𝐷𝜎subscript𝑘𝐵𝑇superscriptsubscript𝐹𝜎𝑖𝑡Δ𝑡subscript𝜉𝜎otherwise\begin{cases}\mathbf{r}_{i}(t+\Delta t)=\mathbf{r}_{i}(t)+\frac{D_{\textmd{T}}% }{k_{B}T}\mathbf{F}_{\textmd{T}}^{(i)}(t)\Delta t+{\bm{\xi}}_{\textmd{T}}\\ \sigma_{i}(t+\Delta t)=\sigma_{i}(t)+\frac{D_{\sigma}}{k_{B}T}F_{\sigma}^{(i)}% (t)\Delta t+\xi_{\sigma}\end{cases}{ start_ROW start_CELL bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t + roman_Δ italic_t ) = bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) + divide start_ARG italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG bold_F start_POSTSUBSCRIPT T end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_t ) roman_Δ italic_t + bold_italic_ξ start_POSTSUBSCRIPT T end_POSTSUBSCRIPT end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t + roman_Δ italic_t ) = italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) + divide start_ARG italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T end_ARG italic_F start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_t ) roman_Δ italic_t + italic_ξ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT end_CELL start_CELL end_CELL end_ROW (21)

where ΔtΔ𝑡\Delta troman_Δ italic_t is the simulation time step. While 𝐅T(i)(t)=𝒓iuext(σi,z;t)jiN𝒓iu(𝐫i𝐫j,σi,σj)superscriptsubscript𝐅T𝑖𝑡subscriptbold-∇subscript𝒓𝑖subscript𝑢extsubscript𝜎𝑖𝑧𝑡superscriptsubscript𝑗𝑖𝑁subscriptbold-∇subscript𝒓𝑖𝑢subscript𝐫𝑖subscript𝐫𝑗subscript𝜎𝑖subscript𝜎𝑗\mathbf{F}_{\textmd{T}}^{(i)}(t)=-\bm{\nabla}_{\bm{r}_{i}}u_{\textmd{ext}}(% \sigma_{i},z;t)-\sum_{j\neq i}^{N}\bm{\nabla}_{\bm{r}_{i}}u(\mathbf{r}_{i}-% \mathbf{r}_{j},\sigma_{i},\sigma_{j})bold_F start_POSTSUBSCRIPT T end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_t ) = - bold_∇ start_POSTSUBSCRIPT bold_italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_z ; italic_t ) - ∑ start_POSTSUBSCRIPT italic_j ≠ italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT bold_∇ start_POSTSUBSCRIPT bold_italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is the translational force acting on colloid i𝑖iitalic_i from both the external field and the pairwise interactions with the other colloids, Fσ(i)(t)=σiψ(σi)σiuext(σi,z;t)jiNσiu(𝐫i𝐫j,σi,σj)superscriptsubscript𝐹𝜎𝑖𝑡subscriptsubscript𝜎𝑖𝜓subscript𝜎𝑖subscriptsubscript𝜎𝑖subscript𝑢extsubscript𝜎𝑖𝑧𝑡superscriptsubscript𝑗𝑖𝑁subscriptsubscript𝜎𝑖𝑢subscript𝐫𝑖subscript𝐫𝑗subscript𝜎𝑖subscript𝜎𝑗F_{\sigma}^{(i)}(t)=-\nabla_{\sigma_{i}}\psi(\sigma_{i})-\nabla_{\sigma_{i}}u_% {\textmd{ext}}(\sigma_{i},z;t)-\sum_{j\neq i}^{N}\nabla_{\sigma_{i}}u(\mathbf{% r}_{i}-\mathbf{r}_{j},\sigma_{i},\sigma_{j})italic_F start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ( italic_t ) = - ∇ start_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_ψ ( italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - ∇ start_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_z ; italic_t ) - ∑ start_POSTSUBSCRIPT italic_j ≠ italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ∇ start_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_u ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is the property force acting on colloid i𝑖iitalic_i resulting from its own free-energy landscape ψ(σ)𝜓𝜎\psi(\sigma)italic_ψ ( italic_σ ), and its interaction with both the external field and the pairwise interactions with the other colloids. Finally, 𝝃Tsubscript𝝃𝑇\bm{\xi}_{T}bold_italic_ξ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and ξσsubscript𝜉𝜎\xi_{\sigma}italic_ξ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT are random vector and random scalar, respectively. Both 𝝃Tsubscript𝝃𝑇\bm{\xi}_{T}bold_italic_ξ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT and ξσsubscript𝜉𝜎\xi_{\sigma}italic_ξ start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT are drawn from a normal distribution with zero mean and variance ξT,αξT,β=2DTΔtδαβdelimited-⟨⟩subscript𝜉𝑇𝛼subscript𝜉𝑇𝛽2subscript𝐷TΔ𝑡subscript𝛿𝛼𝛽\langle\xi_{T,\alpha}\xi_{T,\beta}\rangle=2D_{\textmd{T}}\Delta t\delta_{% \alpha\beta}⟨ italic_ξ start_POSTSUBSCRIPT italic_T , italic_α end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_T , italic_β end_POSTSUBSCRIPT ⟩ = 2 italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT roman_Δ italic_t italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT and ξσ2=2DσΔtdelimited-⟨⟩subscriptsuperscript𝜉2𝜎2subscript𝐷𝜎Δ𝑡\langle\xi^{2}_{\sigma}\rangle=2D_{\sigma}\Delta t⟨ italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ⟩ = 2 italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT roman_Δ italic_t, respectively, with α𝛼\alphaitalic_α, β=x,y,z𝛽𝑥𝑦𝑧\beta=x,y,zitalic_β = italic_x , italic_y , italic_z being the Cartesian components, and δαβsubscript𝛿𝛼𝛽\delta_{\alpha\beta}italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT being the Kronecker delta function.

Simulations were performed using the same geometry and potential parameters as the ones used in RC-DDFT. To avoid non vanishing, non-physical negative values, and extremely large values of the colloid size, the range of σ𝜎\sigmaitalic_σ was limited to [0.1,2]σ00.12subscript𝜎0[0.1,2]\sigma_{0}[ 0.1 , 2 ] italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT by simply rejecting (Monte-Carlo like) moves that would lead to σ<0.1σ0𝜎0.1subscript𝜎0\sigma<0.1\sigma_{0}italic_σ < 0.1 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT or σ>2σ0𝜎2subscript𝜎0\sigma>2\sigma_{0}italic_σ > 2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The same procedure was used to respect the hard walls. Starting with an equilibrium configuration made of 432 responsive colloids (see Fig. 1(a)), the system was perturbed by either switching on or off ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT or ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT, prior to a relaxation run of 30τB30subscript𝜏B30\tau_{\text{B}}30 italic_τ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT during which the system reached a new equilibrium state (see Fig. 1(c)). To obtain the time-dependent density, ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t ), and mean size, σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩, Nrun=4000subscript𝑁𝑟𝑢𝑛4000N_{run}=4000italic_N start_POSTSUBSCRIPT italic_r italic_u italic_n end_POSTSUBSCRIPT = 4000 independent runs were performed. The density profile and the mean size were then averaged over all configurations for a fixed time t𝑡titalic_t.

IV Results and discussion

For our non-equilibrium relaxation study, our methodology consists on applying an external potential,

uext(z;σ;t)=uwalls(z)+ϕ(z,σ;t),subscript𝑢ext𝑧𝜎𝑡subscript𝑢walls𝑧italic-ϕ𝑧𝜎𝑡u_{\text{ext}}(z;\sigma;t)=u_{\text{walls}}(z)+\phi(z,\sigma;t),italic_u start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( italic_z ; italic_σ ; italic_t ) = italic_u start_POSTSUBSCRIPT walls end_POSTSUBSCRIPT ( italic_z ) + italic_ϕ ( italic_z , italic_σ ; italic_t ) , (22)

where uwallssubscript𝑢wallsu_{\text{walls}}italic_u start_POSTSUBSCRIPT walls end_POSTSUBSCRIPT accounts for the confinement effects, and ϕitalic-ϕ\phiitalic_ϕ represents additional external potentials such as osmotic or gravitational potentials, defined further below. In particular, the wall potential is

uwalls(z)={0for0zLforz<0orz>Lsubscript𝑢walls𝑧cases0for0𝑧𝐿otherwiseformulae-sequencefor𝑧0or𝑧𝐿otherwiseu_{\text{walls}}(z)=\begin{cases}0\quad\ \text{for}\quad 0\leq z\leq L\\ \infty\quad\text{for}\quad z<0\ \ \textmd{or}\ \ z>L\end{cases}italic_u start_POSTSUBSCRIPT walls end_POSTSUBSCRIPT ( italic_z ) = { start_ROW start_CELL 0 for 0 ≤ italic_z ≤ italic_L end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL ∞ for italic_z < 0 or italic_z > italic_L end_CELL start_CELL end_CELL end_ROW

We initiate our system at equilibrium and consider two time-dependent protocols. In the first one, the external potential ϕitalic-ϕ\phiitalic_ϕ is absent for t<0𝑡0t<0italic_t < 0, and is activated for t>0𝑡0t>0italic_t > 0 (switch on) to observe the ensuing relaxation towards the new equilibrium stated (reached in the limit t𝑡t\rightarrow\inftyitalic_t → ∞) (cf. Fig. 1 as a representative illustration). In the second protocol we follow the inverse process: the field is already activated for t<0𝑡0t<0italic_t < 0, and it is switched off for t>0𝑡0t>0italic_t > 0. Comparison between both procedures will allow to examine whether the system shows a different relaxation dynamics during the switching on and switching off processes.

This bidirectional exploration is conducted for two external potentials, namely gravitational (ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT) and osmotic (ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT) potentials:

(i) The gravitational external potential is linear in ’height’ z𝑧zitalic_z and reads as

ϕG(z)=Azfor0zL,formulae-sequencesubscriptitalic-ϕG𝑧𝐴𝑧for0𝑧𝐿\phi_{\textmd{G}}(z)=Az\quad\text{for}\quad 0\leq z\leq L,italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT ( italic_z ) = italic_A italic_z for 0 ≤ italic_z ≤ italic_L , (23)

with βAσ0=1𝛽𝐴subscript𝜎01\beta A\sigma_{0}=1italic_β italic_A italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1. Note that the gravitational potential is just a function of position z𝑧zitalic_z, not of size. Still, density inhomogeneities will also affect the compressible particle sizes and their distributions in time and space.

(ii) The osmotic potential, ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ) is given by

ϕO(z,σ)=Bzσ3for0zL,formulae-sequencesubscriptitalic-ϕO𝑧𝜎𝐵𝑧superscript𝜎3for0𝑧𝐿\phi_{\textmd{O}}(z,\sigma)=Bz\sigma^{3}\quad\text{for}\quad 0\leq z\leq L,italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ) = italic_B italic_z italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT for 0 ≤ italic_z ≤ italic_L , (24)

where βBσ04=1𝛽𝐵superscriptsubscript𝜎041\beta B\sigma_{0}^{4}=1italic_β italic_B italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT = 1. This equation introduces a dependence on particle size through the σ3superscript𝜎3\sigma^{3}italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT volume term, aiming to replicate the volumetric effects exerted by osmotic pressure in environments with varying cosolute concentrations with a constant concentration gradient. As such, σ3superscript𝜎3\sigma^{3}italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT represents the volume exclusion effect of particles within this gradient. Denton and Schmidt (2002); Lin, Rotenberg, and Dzubiella (2020)

Refer to caption
Figure 1: Representative snapshots of system configurations made of responsive colloids obtained at different times from Brownian dynamics simulations. For t<0𝑡0t<0italic_t < 0, the system is in equilibrium with hard walls on the left and on the right (solid line). At t=0𝑡0t=0italic_t = 0, the system is perturbed by switching on ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT (solid line) and relaxes (for t>0𝑡0t>0italic_t > 0) before reaching the new equilibrium at t𝑡t\to\inftyitalic_t → ∞. The color gradient from blue (small sizes) to red (large sizes) visualizes the magnitude of the sizes of the colloids.

To validate the theoretical predictions of our extended DDFT, they are compared to BD simulations results. Through this comparison, we analyze macroscopic quantities such as the mean position z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩, mean size σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩, and the pressure exerted on the hard walls. Moreover, we delve into a parametric study focusing on the dynamics of these responsive systems, governed by the parameter α𝛼\alphaitalic_α. For α<1𝛼1\alpha<1italic_α < 1 the swelling/deswelling of the responsive colloid is slower than the translational diffusion, which means that the change of particle size happens later than the structural relation. The opposite occurs for α>1𝛼1\alpha>1italic_α > 1. This dissimilar time relaxation is expected to lead to transient dynamic states, that will be explored in the following section.

Fig. 1 provides a representative illustration of the activation protocol for the particular case of the gravitational external potential. Big and small particles are depicted as red and blue spheres, respectively. The rest of intermediate sizes are represents by a continuous graduation of colors between red and blue. At time t<0𝑡0t<0italic_t < 0 (left panel) the responsive system is in the equilibrium state, confined between two planar hard walls. After activation of the external field (central panel) at t=0𝑡0t=0italic_t = 0, particles dynamically relax in the new potential field and migrate towards the left to lower the external potential energy. Their sizes are also affected subjected to the changes in the local particle concentration. For t𝑡t\rightarrow\inftyitalic_t → ∞ (right panel) the system finally reaches the final equilibrium state in presence of the external potential, with a very different profile and size distribution compared to the original state.

IV.1 Gravitational potential

Refer to caption
Figure 2: Time evolution of the density ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t ) and size σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ of the RCs in the gravitational field, ϕG(z)subscriptitalic-ϕG𝑧\phi_{\textmd{G}}(z)italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT ( italic_z ), plotted at different times ranging from t=0𝑡0t=0italic_t = 0 (dark blue) to 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (yellow), with a time interval of Δt=0.5τBΔ𝑡0.5subscript𝜏B\Delta t=0.5\tau_{\rm B}roman_Δ italic_t = 0.5 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Lines represent the theoretical predictions obtained with RC-DDFT, whereas symbols correspond to BD simulations. Panels (a) and (c) show, respectively, the local density (ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t )) and the local mean size (σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\left\langle\sigma(z;t)\right\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩) obtained when the external potential is switched on for t>0𝑡0t>0italic_t > 0. Similarly, panels (b) and (d) show the same quantities, but for the deactivation process (’switched off’). In this case, σ(z)delimited-⟨⟩𝜎𝑧\left\langle\sigma(z)\right\rangle⟨ italic_σ ( italic_z ) ⟩ data of RC-DDFT is multiplied by a factor 1.051.051.051.05 for a clearer comparison. All calculations are performed considering α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and a surface density N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

We start analyzing the relaxation dynamics of a responsive colloidal system immersed in the gravitational external potential given by Eq. (23) under the two protocols (switching on and off) outlined in the preceding section. For this case, we select α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. Fig. 2 depicts the time evolution of the mean local density (ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t )) and local mean size (σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩) from t=0𝑡0t=0italic_t = 0 to 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT at Δt=0.5τBΔ𝑡0.5subscript𝜏B\Delta t=0.5\tau_{\rm B}roman_Δ italic_t = 0.5 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT intervals. Panels (a) and (c) show the results following the activation (’switch on’) of the gravitational field, while panels (b) and (d) depict the outcomes subsequent to its deactivation (’switch off’) .

The analysis of ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ), depicted in panels (a) and (b), reveals a significant alignment between the BD and RC-DDFT methods throughout the dynamic process. These panels also indicate that the system has reached nearly the final equilibrium state already for t3τB𝑡3subscript𝜏Bt\approx 3\tau_{\rm B}italic_t ≈ 3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Conversely, the evaluation of σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ in panels (c) and (d) demonstrates consistent conformity between the methods, albeit with notable differences, particularly a consistent disparity of approximately 5% throughout the dynamic sequence. This deviation in σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ arises not only in this confined geometry but also in bulk systems (without external potentials and walls). We attribute this discrepancy to the inherent limitations of the mean-field approximation employed in RC-DDFT. To facilitate clearer comparison in the figures, this difference has been rectified in the RC-DDFT data by a scaling factor of 1.051.051.051.05. Following this adjustment, the temporal evolution described by both techniques closely corresponds.

In the initial stage (t=0𝑡0t=0italic_t = 0) of the activation state, depicted in panel (a) of Fig. 2, we observe a nearly flat density profile, ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ), with the exception of regions near the walls where particle adsorption effects become prominent. This accumulation near the hard walls occurs because particles are pushed from the bulk to the walls due to the interparticle repulsive interactions. In addition, the local mean size, σ(z)delimited-⟨⟩𝜎𝑧\langle\sigma(z)\rangle⟨ italic_σ ( italic_z ) ⟩, as shown in Fig.2(c), exhibits intriguing behavior near the wall. Typically, in the bulk, the mean size is expected to decrease as ρ𝜌\rhoitalic_ρ increases. Baul and Dzubiella (2021) However, near the walls, an increase in σdelimited-⟨⟩𝜎\langle\sigma\rangle⟨ italic_σ ⟩ is observed as ρ𝜌\rhoitalic_ρ increases, a phenomenon also noted in equilibrium configurations. Moncho-Jordá, Groh, and Dzubiella (2024) This behavior can be explained by the conditions faced by colloids near a hard boundary. Unlike in the bulk, where colloids are fully surrounded by other particles, being near a wall reduces the number of neighboring particles by half. This reduction in surrounding particles diminishes the overall repulsive forces acting on the colloids, allowing them to expand.

Upon activation of the gravitational field ϕG(z)subscriptitalic-ϕG𝑧\phi_{\textmd{G}}(z)italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT ( italic_z ) at time t>0𝑡0t>0italic_t > 0, the system is subjected to a reordering as particles migrate towards the region with lower external field, located on the left side of the slit. This process, depicted in the dynamic sequence in Figs. 2(a) and (c) is not instantaneous; thus, we plot the density and mean size at various times to capture the evolution. It is interesting to remark the strong change of σ(z)delimited-⟨⟩𝜎𝑧\langle\sigma(z)\rangle⟨ italic_σ ( italic_z ) ⟩ during the first stages of the evolution (0.5τB0.5subscript𝜏B0.5\tau_{\rm B}0.5 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT), compared to the evolution of ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t ). Given that α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 implies that size diffusion should be significantly slower than spatial diffusion, the observed rapid change in size must be necessarily caused by Stokes-type diffusion, DT1/σsimilar-tosubscript𝐷T1𝜎D_{\textmd{T}}\sim 1/\sigmaitalic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT ∼ 1 / italic_σ: Smaller particles, having a higher diffusion coefficient, move faster towards the left, leaving behind larger particles and thus increasing the average size on the right side of the system. Additionally, a decrease in density favors more expanded states for the particles.

The system keeps evolving, and we illustrate these dynamics only up to 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT because the changes become exceedingly slow thereafter. By 10τB10subscript𝜏B10\tau_{\rm B}10 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT, the system reaches a state indistinguishable from the equilibrium state achieved under the gravitational field. In this final equilibrium state, which is also the starting point for the deactivation process (Figs. 2(b) and (d)), the density ρ𝜌\rhoitalic_ρ decays from the left wall. In an ideal, non-interacting system, this decay would be purely exponential. However, due to the existent repulsive interactions between colloids, the actual final equilibrium density profile departs from this ideal behavior. Excluding the near-wall effects previously discussed, regions with higher particle concentration correspond to smaller mean sizes, indicating compression due to increased density. Interestingly, near the right wall, where ρ𝜌\rhoitalic_ρ is nearly zero, the mean size is maximal and very close to one, suggesting negligible interparticle interactions as the size distribution aligns with the parent distribution, p(σ)𝑝𝜎p(\sigma)italic_p ( italic_σ ). The complete dynamics of the activation process depicted in Fig. 1 are illustrated through representative snapshots obtained from Brownian Dynamics (BD) simulations, capturing the initial, intermediate, and final states. As depicted, upon activation of ϕG(z)subscriptitalic-ϕG𝑧\phi_{\textmd{G}}(z)italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT ( italic_z ), responsive particles undergo noticeable diffusion towards the left wall. This results in compression, leading to a significant localization effect in particle size when compared to the less compressed region near the right wall.

Refer to caption
Figure 3: Time evolution of the RCs in the osmotic field, ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ), plotted at different times ranging from t=0𝑡0t=0italic_t = 0 (dark blue) to 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (yellow), with a time interval of Δt=0.5τBΔ𝑡0.5subscript𝜏B\Delta t=0.5\tau_{\rm B}roman_Δ italic_t = 0.5 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Lines represent the theoretical predictions obtained with RC-DDFT, whereas symbols correspond to BD simulations. Panels (a) and (c) show respectively the local density (ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t )) and the local mean size (σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\left\langle\sigma(z;t)\right\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩) obtained when the external potential is switched on for t>0𝑡0t>0italic_t > 0. Similarly, panels (b) and (d) show the same quantities, but for the deactivation process. All calculations are performed considering α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and a surface density N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

Figs. 2(b) and (d) show again ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t ) and σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ for the deactivation process, respectively. In this case, we find that the migration towards the new equilibrium state is not symmetrical compared to the activation process. During activation, particle movement is driven by the external gravitational force, whereas upon deactivation, particle movement is driven by the gradient of concentration, which pushes the particle from the more dense region (left) to the more dilute one (right). This asymmetry results in differing dynamics between activation and deactivation phases, with the gravitational field’s application appearing to accelerate the dynamic process as it will be discussed later.

IV.2 Osmotic potential

Next, we turn to examine the dynamics of the RC fluid when the osmotic potential is activated and deactivated (Eq. 24). This external potential varies linearly with z𝑧zitalic_z, akin to gravitational force, but its pronounced σ3superscript𝜎3\sigma^{3}italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT-dependency gives rise to a markedly distinct qualitative dynamic behavior. Fig. 3(a) and (c) illustrate the mean particle density and mean size within the planar slit during the activation process, respectively, while plots (b) and (d) delineate the same quantities during deactivation. We maintain α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

Upon activation of ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ), particles tend to migrate globally towards the left, mitigating the energy contribution induced by the external potential, expressed as ρϕO𝑑z𝑑σ𝜌subscriptitalic-ϕOdifferential-d𝑧differential-d𝜎\int\rho\phi_{\textmd{O}}dzd\sigma∫ italic_ρ italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT italic_d italic_z italic_d italic_σ. Analogous to gravitational effects, this results in a notable increase in particle density near the left wall over time.

However, σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ exhibits contrasting behavior, with larger particle sizes accompanying high-density regions, as depicted in Fig. 3(c). This occurs despite the compression near the left wall. The phenomenon can be elucidated by the strong dependence of ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ) on σ𝜎\sigmaitalic_σ: the energetic penalty for larger colloids near the right wall prompts smaller colloids to prevail in that region. Additionally, σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ experiences rapid decline in early stages of evolution (t<0.5τB𝑡0.5subscript𝜏Bt<0.5\tau_{\textmd{B}}italic_t < 0.5 italic_τ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT), not attributable to RC shrinkage, given that calculations are conducted with a deswelling diffusion 10101010 times smaller than spatial diffusion. This effect stems from the migration of larger colloids towards the left wall, propelled by the applied external force, fext=dϕO/dz=Aσ3subscript𝑓ext𝑑subscriptitalic-ϕO𝑑𝑧𝐴superscript𝜎3f_{\textmd{ext}}=-d\phi_{\textmd{O}}/dz=-A\sigma^{3}italic_f start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT = - italic_d italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT / italic_d italic_z = - italic_A italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

Distinct dynamical behavior emerges upon deactivating the osmotic potential. Remarkably, convergence is achieved within 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT in both scenarios. However, activating the osmotic potential leads to markedly faster dynamics compared to its deactivation. This acceleration is particularly pronounced when examining the size distribution, with temporal profiles nearly coinciding after just tτB𝑡subscript𝜏Bt\approx\tau_{\rm B}italic_t ≈ italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT.

Furthermore, transient behavior is observed upon activation of ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ). Indeed, after activation, there is a rapid migration of colloids towards the left wall, resulting in a significant concentration increase at very short times in this region. Concurrently, the mean particle size of colloids near the left wall initially increases, followed by a decay over longer times. This clearly indicates that particles first move to the left and then decreases their size.

Conversely, deactivating ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ) yields the opposite trend, with ρ𝜌\rhoitalic_ρ and σ𝜎\sigmaitalic_σ increasing on the left and decreasing on the right until equilibrium is attained.

In both scenarios (activation and deactivation), the comparison between theoretical predictions from RC-DDFT and BD simulations demonstrates very good agreement.

IV.3 Macroscopic Analysis

Refer to caption
Figure 4: Time evolution of (a) the center of mass position (z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩), (b) the mean particle size of the system (σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩, with gravitational RC-DDFT data multiplied by 1.05), (c) the pressure applied by the fluid on the left wall (Pleft(t)subscript𝑃left𝑡P_{\textmd{left}}(t)italic_P start_POSTSUBSCRIPT left end_POSTSUBSCRIPT ( italic_t )), and (d) on the right wall Pright(t)subscript𝑃𝑟𝑖𝑔𝑡𝑡P_{right}(t)italic_P start_POSTSUBSCRIPT italic_r italic_i italic_g italic_h italic_t end_POSTSUBSCRIPT ( italic_t ), obtained for the activation (solid lines) and deactivation (dashed lines) processes. Orange and blue colors represents the results for the gravitational and osmotic potentials, respectively. Lines denote RC-DDFT predictions, whereas symbols BD simulations. In all cases α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and a surface density N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.
Refer to caption
Figure 5: RC-DDFT results for the time evolution of the system structure for the osmotic field, ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT, being switched on at time t=0𝑡0t=0italic_t = 0 for various values of the timescale parameter α𝛼\alphaitalic_α. (a), (b) Local density, ρ(z)𝜌𝑧\rho(z)italic_ρ ( italic_z ), and (c), (d) local mean size, σ(z)delimited-⟨⟩𝜎𝑧\left\langle\sigma(z)\right\rangle⟨ italic_σ ( italic_z ) ⟩, at different times ranging from 0 τBsubscript𝜏B\tau_{\rm B}italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (blue) to 3 τBsubscript𝜏B\tau_{\rm B}italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT (yellow) with a time interval of 0.5 τBsubscript𝜏B\tau_{\rm B}italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. The surface density is N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. The insert in (b) shows the time evolution of the density ρ𝜌\rhoitalic_ρ and the mean size σdelimited-⟨⟩𝜎\left\langle\sigma\right\rangle⟨ italic_σ ⟩ on the left wall.

To validate the agreement between RC-DDFT and BD beyond the local density and size profiles, we examine three interesting macroscopic, integrated quantities. These quantities not only offer insights into the global behavior of the system but also facilitate a comparison of the time scales on which dynamical processes occur. The quantities of interest are the center of mass position zdelimited-⟨⟩𝑧\langle z\rangle⟨ italic_z ⟩ of the suspension (Eq. (17)), the global mean particle size σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩ (Eq. (19)), and the pressures exerted by the RC fluid on the left and right hard walls, given by Pleft(t)subscript𝑃left𝑡P_{\textmd{left}}(t)italic_P start_POSTSUBSCRIPT left end_POSTSUBSCRIPT ( italic_t ) and Pright(t)subscript𝑃right𝑡P_{\textmd{right}}(t)italic_P start_POSTSUBSCRIPT right end_POSTSUBSCRIPT ( italic_t ) (Eq. (20)), respectively.

The time evolution of these quantities is illustrated in Fig. 4 for α=0.1𝛼0.1\alpha=0.1italic_α = 0.1 and N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. Each panel presents results obtained from RC-DDFT for activated (solid lines) and deactivated (dashed lines) gravitational and osmotic external potentials, distinguished by orange and blue lines, respectively. Corresponding BD simulations are represented by squared and circled symbols. Remarkable congruence between macroscopic quantities derived from both RC-DDFT and BD, not only in shape but also in indicative time scales (the values of DDFT-σ𝜎\sigmaitalic_σ under gravitational field has been multiplied again by a factor 1.05).

The analysis of these curves reveals several features of the relaxation process. Primarily, it is evident that relaxation during the activation of the external field is consistently faster than during deactivation. We attribute this behavior to the supplementary driving force induced by the activated external potential, intensifying the motion of confined colloids. Conversely, during the deactivation process, this external force is absent, resulting in colloids relying solely on diffusion for movement, thereby leading to a slower evolution towards equilibrium.

With our system’s inherent two degrees of freedom, we employ a double exponential function to fit these macroscopic quantities, facilitating the extraction of time scales:

ψ(t)=A1et/τ1+A2et/τ2+c,𝜓𝑡subscript𝐴1superscript𝑒𝑡subscript𝜏1subscript𝐴2superscript𝑒𝑡subscript𝜏2𝑐\psi(t)=A_{1}e^{-t/\tau_{1}}+A_{2}e^{-t/\tau_{2}}+c,italic_ψ ( italic_t ) = italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_c , (25)

being ψ(t)={z(t),σ(t),PL(t),PL(t)}𝜓𝑡delimited-⟨⟩𝑧𝑡delimited-⟨⟩𝜎𝑡subscript𝑃𝐿𝑡subscript𝑃𝐿𝑡\psi(t)=\{\langle z(t)\rangle,\langle\sigma(t)\rangle,P_{L}(t),P_{L}(t)\}italic_ψ ( italic_t ) = { ⟨ italic_z ( italic_t ) ⟩ , ⟨ italic_σ ( italic_t ) ⟩ , italic_P start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_t ) , italic_P start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( italic_t ) }. The relaxation times, τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, obtained from these fits, are summarized in Table 1, presenting a comparison of the time scales predicted by RC-DDFT and observed through BD (between parenthesis) under the scenarios studied.

Table 1: Relaxation times obtained by fitting Eq. (25) to mean size, σ(t)delimited-⟨⟩𝜎𝑡\left\langle\sigma(t)\right\rangle⟨ italic_σ ( italic_t ) ⟩, and mean position, z(t)delimited-⟨⟩𝑧𝑡\left\langle z(t)\right\rangle⟨ italic_z ( italic_t ) ⟩ predicted by RC-DDFT during the activation (on) and deactivation (off) of the external potentials, ϕG(z)subscriptitalic-ϕG𝑧\phi_{\textmd{G}}(z)italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT ( italic_z ) and ϕO(z,σ)subscriptitalic-ϕO𝑧𝜎\phi_{\textmd{O}}(z,\sigma)italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT ( italic_z , italic_σ ). The corresponding times for BD simulations are shown inside the parenthesis.
τ𝜏\tauitalic_τ ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT (on) ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT (off) ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT (on) ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT (off)
zdelimited-⟨⟩𝑧\langle z\rangle⟨ italic_z ⟩ τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 0.90 (0.91) 1.38 (1.47) 0.24 (0.25) 1.33 (1.46)
τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 1.97 (1.95) 1.38 (1.47) 1.23 (1.23) 1.33 (1.46)
σdelimited-⟨⟩𝜎\langle\sigma\rangle⟨ italic_σ ⟩ τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT 1.08 (1.15) 0.61 (0.72) 0.30 (0.31) 0.28 (0.27)
τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT 1.69 (1.83) 1.50 (1.55) 1.20 (1.42) 1.48 (1.61)

In every scenario, we observe two distinct time scales that are similar within each potential and across both RC-DDFT and BD analyses. This similarity suggests that the complex dynamics of the system are influenced by the interaction between translational and size diffusion processes. The differences in the shapes of macroscopic quantities for various potentials, as shown in Fig. 4, likely arise from the differing impacts of these time scales. The close match between macroscopic quantities obtained from BD simulations and those predicted by RC-DDFT highlights the effectiveness of the RC-DDFT extension in accurately capturing the dynamics of the system. This concordance further supports the use of RC-DDFT for additional studies.

As summarized in Table 1, the relaxation time for z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩ when ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT is switched on (0.24τB0.24subscript𝜏B0.24\tau_{\rm B}0.24 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and off (1.33τB1.33subscript𝜏B1.33\tau_{\rm B}1.33 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) or the relaxation time for σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩ when ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT is switched on (1.08τB1.08subscript𝜏B1.08\tau_{\rm B}1.08 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) and off (0.61τB0.61subscript𝜏B0.61\tau_{\rm B}0.61 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT) is a clear signature of irreversible relaxation pathway. In addition, the difference in relaxation time for z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩ when switching on ϕOsubscriptitalic-ϕO\phi_{\textmd{O}}italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT and ϕGsubscriptitalic-ϕG\phi_{\textmd{G}}italic_ϕ start_POSTSUBSCRIPT G end_POSTSUBSCRIPT reveals that the nature of the perturbation has a important effect on the relaxation process (see Fig. 4).

IV.4 Competition of time scales: the α𝛼\alphaitalic_α-study

Having evaluated the effects of the activation and deactivation process of the external field, we now focus on the assessment of the influence of the time scale ratio, α=Dσ/D0𝛼subscript𝐷𝜎subscript𝐷0\alpha=D_{\sigma}/D_{0}italic_α = italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT / italic_D start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, on the non-equilibrium relaxation dynamics of our confined system of soft RCs. As mentioned earlier, α𝛼\alphaitalic_α modulates the interplay between size diffusion and translational diffusion. For such a study, we select the osmotic external potential and consider only the activation process. Fig. 5 shows the time evolution of ρ(z;t)𝜌𝑧𝑡\rho(z;t)italic_ρ ( italic_z ; italic_t ) and σ(z;t)delimited-⟨⟩𝜎𝑧𝑡\langle\sigma(z;t)\rangle⟨ italic_σ ( italic_z ; italic_t ) ⟩ for α=1𝛼1\alpha=1italic_α = 1 (plots (a) and (c)) and α=0.01𝛼0.01\alpha=0.01italic_α = 0.01 (plots (b) and (d)), representing two limiting dynamic condition for which the σ𝜎\sigmaitalic_σ-diffusion is very fast and very slow compared to the translational diffusion. The particle surface concentration is fixed in all cases to N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. Considering the close alignment between RC-DDFT and BD observed in previous sections, we only depict the results obtained with RC-DDFT.

For α=1𝛼1\alpha=1italic_α = 1, the particle size rapidly adjusts to the new environmental conditions during the initial relaxation stage, and the system approaches full structural equilibrium within 3τB3subscript𝜏B3\tau_{\rm B}3 italic_τ start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT. Decreasing α𝛼\alphaitalic_α to 0.010.010.010.01 results in a slower response of particle size to the activation of the external potential. In this regime, our findings reveal the presence of a transient dynamic state: initially, particles accumulate on the left wall for t<3τB𝑡3subscript𝜏Bt<3\tau_{\textmd{B}}italic_t < 3 italic_τ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT, followed by a subsequent decrease in concentration towards the final equilibrium state for t>3τB𝑡3subscript𝜏Bt>3\tau_{\textmd{B}}italic_t > 3 italic_τ start_POSTSUBSCRIPT B end_POSTSUBSCRIPT. This dynamic reentrance is clearly depicted in the inset of Fig. 5(b), illustrating the time dependence of the particle density in contact with the left hard wall, ρ(z=0,t)𝜌𝑧0𝑡\rho(z=0,t)italic_ρ ( italic_z = 0 , italic_t ), which exhibits a maximum at t3τB𝑡3subscript𝜏𝐵t\approx 3\tau_{B}italic_t ≈ 3 italic_τ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT. A similar trend is observed for the mean size of colloids in contact with the left wall (σ(z=0,t)delimited-⟨⟩𝜎𝑧0𝑡\langle\sigma(z=0,t)\rangle⟨ italic_σ ( italic_z = 0 , italic_t ) ⟩), initially increasing with time before eventually decreasing again (see again the inset of Fig. 5(b)).

It is important to highlight that this transient dynamic state observed at intermediate times arises directly from the disparity in time scales between translation and swelling, evident for α=0.01𝛼0.01\alpha=0.01italic_α = 0.01 and absent for α=1𝛼1\alpha=1italic_α = 1. Initially, colloids diffuse towards the left under the external field without altering their size, which evolves 100100100100 times slower. The observed increase in σ(z=0,t)delimited-⟨⟩𝜎𝑧0𝑡\langle\sigma(z=0,t)\rangle⟨ italic_σ ( italic_z = 0 , italic_t ) ⟩ is primarily attributed to the selective motion of larger colloids driven by the applied external force fext=dϕO/dz=Aσ3subscript𝑓ext𝑑subscriptitalic-ϕO𝑑𝑧𝐴superscript𝜎3f_{\textmd{ext}}=-d\phi_{\textmd{O}}/dz=-A\sigma^{3}italic_f start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT = - italic_d italic_ϕ start_POSTSUBSCRIPT O end_POSTSUBSCRIPT / italic_d italic_z = - italic_A italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. Over longer timescales, compression effects lead to a gradual (slower) reduction in particle size, consequently inducing a decrease in ρ(z=0,t)𝜌𝑧0𝑡\rho(z=0,t)italic_ρ ( italic_z = 0 , italic_t ) until the new equilibrium state is attained. In essence, this observation signifies a transition from translationally driven to size-driven relaxation dynamics as α𝛼\alphaitalic_α decreases from 1111 to 0.010.010.010.01, a characteristic exclusive to responsive colloids.

Refer to caption
Figure 6: Time evolution of (a) the center of mass position (z(t)delimited-⟨⟩𝑧𝑡\left\langle z(t)\right\rangle⟨ italic_z ( italic_t ) ⟩) and (b) the mean size of the colloids (σ(t)delimited-⟨⟩𝜎𝑡\left\langle\sigma(t)\right\rangle⟨ italic_σ ( italic_t ) ⟩) predicted by RC-DDFT (symbols) during the activation of the osmotic external potential, obtained for different values of α𝛼\alphaitalic_α from 0.010.010.010.01 to 2222. Solid lines are the result of the fitting of the data using Eq. (25). In all cases N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

This rich interplay between structural and size relaxation is further illustrated in Fig. 6(a) and (b), which explore, respectively, the impact of α𝛼\alphaitalic_α on the center of mass position (z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩) and on the mean particle size (σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩) across a wide range of α𝛼\alphaitalic_α values from 0.010.010.010.01 to 2222. In cases where α<0.04𝛼0.04\alpha<0.04italic_α < 0.04, a non-monotonic curve with a pronounced minimum appears in z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩, signifying that responsive colloids initially migrate towards the left wall due to the applied external force before subsequently reversing direction due to gradual size reduction. Conversely, for α>0.25𝛼0.25\alpha>0.25italic_α > 0.25, the opposing trend is found: the particle size undergoes rapid reduction attributed to the compression induced by the osmotic field during the initial stages of evolution, yet the colloids have not traversed the necessary distance to achieve structural relaxation. For longer times, the progressive redistribution of colloids towards the left wall through spatial diffusion allows σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩ to exhibit a slight increase, driven by the swelling of colloids near z=0𝑧0z=0italic_z = 0.

This transient behavior observed for dissimilar values of DTsubscript𝐷TD_{\textmd{T}}italic_D start_POSTSUBSCRIPT T end_POSTSUBSCRIPT and Dσsubscript𝐷𝜎D_{\sigma}italic_D start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT becomes evident in Fig. 7, where z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩ is plotted against σ(t)delimited-⟨⟩𝜎𝑡\langle\sigma(t)\rangle⟨ italic_σ ( italic_t ) ⟩ for different values of α𝛼\alphaitalic_α, from 0.010.010.010.01 to 3333. Indeed, for α=0.01𝛼0.01\alpha=0.01italic_α = 0.01 we find that z(t)delimited-⟨⟩𝑧𝑡\langle z(t)\rangle⟨ italic_z ( italic_t ) ⟩ shows a dynamic reentrance induced by the slow σ𝜎\sigmaitalic_σ-response, leading to a minimum in the curve. Conversely, for α=3𝛼3\alpha=3italic_α = 3. the curve depicts a shoulder on the left, indicating the reentrance of the center of mass location caused by the slow translational diffusion. These phenomena controlled by α𝛼\alphaitalic_α demonstrate the dynamic competition between translational and size diffusion. At very low or high α𝛼\alphaitalic_α values, the dynamics of mean size and position appear closely coupled, depicting a scenario where one aspect of the system’s behavior must "wait" for the other to stabilize.

Refer to caption
Figure 7: Transition pathways: Center of mass location zdelimited-⟨⟩𝑧\langle z\rangle⟨ italic_z ⟩ versus the mean particle size σdelimited-⟨⟩𝜎\langle\sigma\rangle⟨ italic_σ ⟩ of the system during the relaxation process after activation of the osmotic external potential, plotted for different values of α𝛼\alphaitalic_α. The red point on the right-top represents the initial equilibrium state (t=0𝑡0t=0italic_t = 0), whereas the one located on the left-bottom corresponds to the new equilibrium state (t𝑡t\rightarrow\inftyitalic_t → ∞). Results are obtained using RC-DDFT with N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT.

The fitting of mean size and mean position according to the double exponential function given by Eq. (25) yields relaxation times and prefactors as functions of α𝛼\alphaitalic_α, as shown in Figure 8. This analysis confirms the presence of two distinct time scales. Examining the mean position, as depicted in Fig. 8(a), enables us to classify τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as the translational relaxation time and τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as the size relaxation time. Notably, τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT remains relatively unchanged across different values of α𝛼\alphaitalic_α. In scenarios where α𝛼\alphaitalic_α is high, the prefactor A1subscript𝐴1A_{1}italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is always significantly larger then A2subscript𝐴2A_{2}italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (please not that in this regime A2subscript𝐴2A_{2}italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is very close to 00.), indicating that size adjusts swiftly, and the mean position is predominantly governed by translational relaxation. Conversely, at lower α𝛼\alphaitalic_α values, the significance of A2subscript𝐴2A_{2}italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT elevates, taking negative values. In this regime the translational relaxation transpires swiftly following a change in size, suggesting that the mean position, zdelimited-⟨⟩𝑧\langle z\rangle⟨ italic_z ⟩, is driven by size adjustments. Taking a look to the relaxation times of the mean size, depicted in Fig. 8(b), we can conversely identify τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT as the size relaxation, since it follows a power law with α𝛼\alphaitalic_α, and τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT as the translational or density relaxation. For high α𝛼\alphaitalic_α the relaxation in size is mainly dominated by size diffusion, since A1A2much-greater-thansubscript𝐴1subscript𝐴2A_{1}\gg A_{2}italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≫ italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

Refer to caption
Refer to caption
Figure 8: Relaxation times, τisubscript𝜏𝑖\tau_{i}italic_τ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and prefactors, Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, as a function of α𝛼\alphaitalic_α. They are obtained by fitting z(t)delimited-⟨⟩𝑧𝑡\left\langle z(t)\right\rangle⟨ italic_z ( italic_t ) ⟩ (plots (a) and (c)) andσ(t)delimited-⟨⟩𝜎𝑡\left\langle\sigma(t)\right\rangle⟨ italic_σ ( italic_t ) ⟩ (plots (b) and (d)) predicted by RC-DDFT to Eq. (25), during the activation process of the osmotic external potential, for N/S=1.2σ02𝑁𝑆1.2superscriptsubscript𝜎02N/S=1.2\sigma_{0}^{-2}italic_N / italic_S = 1.2 italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. The data corresponds to the structural time evolution shown in Figs. 6.

The conspicuous discontinuities observed in Figs. 8(a) and (b) might be interpreted, from a physical perspective, as indicative of a behavioral transition in the dynamical relaxation. Specifically, the domain characterized by high α𝛼\alphaitalic_α values can be conceptualized as a conventional liquid-like phase where translational relaxation predominates as the chief mechanism. Conversely, the domain with low α𝛼\alphaitalic_α values mirrors a more glassy/amorphous phase, with size relaxation emerging as the primary mechanism and a linked localized cage relaxation. This delineation hints at a fundamental shift in the system’s behavior, transitioning from one dominated by translational dynamics to one where size relaxations play a more critical role.

V Concluding remarks

In this work, we have developed and combined a dynamic density-functional theory (RC-DDFT) framework and Brownian dynamics (BD) simulations to investigate the full time-dependent non-equilibrium relaxation dynamics of confined systems of soft responsive colloids (RCs) after activation/deactivation of external potentials. In contrast to conventional models of soft colloids, in our model the size dynamics of the colloids is explicitly resolved and the influence of its relaxation behavior and timescale on the full system could be explored for the first time. The results showed a complex interplay between the translational diffusion and particle swelling/shrinking, leading to interesting non-equilibrium structuring as well as reentrant transient states when the typical spatial and size relaxation times scales are very different. We also demonstrated that the modification of intrinsic timescales of the system lead to tuneable macroscopic relaxation times and pathways in systems with many relaxing degrees of freedom. The excellent agreement between DDFT and BD showed that DDFT can also be faithfully used to study the nonequilibrium behavior of soft, weakly correlated colloids with additional internal degrees of freedom, if the systems are not too slaved by dissipative mechanisms. Schmidt (2022) Hence, our study constitutes a big step forward for the modeling and description of the full nonequilibrium structuring of soft complex fluids.

In future, it would be interesting to extend the RC-DDFT method to responsive (intrinsically polydisperse) systems of stiffer systems, e.g., microgels modeled by Hertizan potentials. Urich and Denton (2016); Rovigatti et al. (2019) For these systems, we expect to find marked oscillations of the density profiles that will become strongly affected by particle stiffness. Also the dynamical properties can lead to the formation of interesting transient, e.g., highly structured while full nonequilibrium regions that finally relax when the colloids adapt their size to the new environmental conditions. These transients may have interesting new properties. Finally, one could also envision to add more internal degrees of freedom to the colloids with a more complex hierarchy of timescales, or even add internal chemical activity to the colloids, e.g., as used in catalytically active nanoreactors where the bistable size response is crucial for complex self-dynamics. Milster et al. (2023)

Acknowledgements.
This work was supported by the Deutsche Forschungsgemeinschaft (DFG) via the Research Unit FOR 5099 “Reducing complexity of nonequilibrium systems” and Project No. 430195928. The authors also acknowledge support by the state of Baden-Württemberg through bwHPC and the DFG through grant no INST 39/963-1 FUGG (bwForCluster NEMO). Further support was provided by grant PID2022-136540NB-I00 funded by MCIN/AEI/10.13039/501100011033, ERDF A way of making Europe, the Spanish Ministerio de Ciencia e Innovación, Programa Estatal de Investigación Científica, Técnica y de Innovación 2021–2023 (Project No. PID2022-136540NB-00) and program Visiting Scholars funded by the Plan Propio of the University of Granada (Project No. PPVS2018-08). J.L.M. thanks the Ph.D. student fellowship (FPU21/03568) supported by the Spanish Ministerio de Universidades.

References

  • Stuart et al. (2010) M. A. C. Stuart, W. T. S. Huck, J. Genzer, M. Müller, C. Ober, M. Stamm, G. B. Sukhorukov, I. Szleifer, V. V. Tsukruk, M. Urban, F. Winnik, S. Zauscher, I. Luzinov,  and S. Minko, “Emerging applications of stimuli-responsive polymer materials,” Nat. Mater. 9, 101 (2010).
  • Onuchic, Luthey-Schulten, and Wolynes (1997) J. N. Onuchic, Z. Luthey-Schulten,  and P. G. Wolynes, “Theory of protein folding: The energy landscape perspective,” Annu. Rev. Phys. Chem. 48, 545–600 (1997).
  • Wu and Wang (1998) C. Wu and X. Wang, “Globule-to-coil transition of a single homopolymer chain in solution,” Phys. Rev. Lett. 80, 4092–4094 (1998).
  • Cho, Levy, and Wolynes (2006) S. S. Cho, Y. Levy,  and P. G. Wolynes, “P versus Q: Structural reaction coordinates capture protein folding on smooth landscapes,” Proc. Natl. Acad. Sci. U.S.A. 103, 586–591 (2006).
  • Choi et al. (2011a) U. B. Choi, J. J. McCann, K. R. Weninger,  and M. E. Bowen, “Beyond the random coil: Stochastic conformational switching in intrinsically disordered proteins,” Structure 19, 566 – 576 (2011a).
  • Denton and Schmidt (2002) A. R. Denton and M. Schmidt, “Colloid-induced polymer compression,” J. Phys.-Condes. Matter 14, 12051–12062 (2002).
  • Urich and Denton (2016) M. Urich and A. R. Denton, “Swelling, structure, and phase stability of compressible microgels,” Soft Matter 12, 9086–9094 (2016).
  • Brijitta and Schurtenberger (2019) J. Brijitta and P. Schurtenberger, “Responsive hydrogel colloids: Structure, interactions, phase behavior, and equilibrium and nonequilibrium transitions of microgel dispersions,” Curr. Opin. Colloid Interface Sci. 40, 87–103 (2019).
  • Baul and Dzubiella (2021) U. Baul and J. Dzubiella, “Structure and dynamics of responsive colloids with dynamical polydispersity,” J. Phys.-Condes. Matter 33, 174002 (2021).
  • Baul et al. (2021a) U. Baul, N. Göth, M. Bley,  and J. Dzubiella, “Modulating internal transition kinetics of responsive macromolecules by collective crowding,” J. Chem. Phys. 155 (2021a).
  • Lin, Rotenberg, and Dzubiella (2020) Y.-C. Lin, B. Rotenberg,  and J. Dzubiella, “Structure and position-dependent properties of inhomogeneous suspensions of responsive colloids,” Phys. Rev. E 102, 042602 (2020).
  • Meng and Li (2013) H. Meng and G. Li, “Reversible switching transitions of stimuli-responsive shape changing polymers,” J. Mater. Chem. A 1, 7838–7865 (2013).
  • Lee et al. (2019) J. Lee, K. H. Ku, C. H. Park, Y. J. Lee, H. Yun,  and B. J. Kim, “Shape and color switchable block copolymer particles by temperature and ph dual responses,” ACS Nano 13, 4230–4237 (2019).
  • Lim and Denton (2014) W. K. Lim and A. R. Denton, “Polymer crowding and shape distributions in polymer-nanoparticle mixtures,” J. Chem. Phys. 141, 114909 (2014).
  • Lim and Denton (2016a) W. K. Lim and A. R. Denton, “Depletion-induced forces and crowding in polymer-nanoparticle mixtures: Role of polymer shape fluctuations and penetrability,” J. Chem. Phys. 144, 024904 (2016a).
  • Lim and Denton (2016b) W. K. Lim and A. R. Denton, “Influence of polymer shape on depletion potentials and crowding in colloid–polymer mixtures,” Soft Matter 12, 2247–2252 (2016b).
  • Harrer et al. (2019) J. Harrer, M. Rey, S. Ciarella, H. Löwen, L. M. C. Janssen,  and N. Vogel, “Stimuli-responsive behavior of pnipam microgels under interfacial confinement,” Langmuir 35, 10512–10521 (2019).
  • Del Monte and Zaccarelli (2024) G. Del Monte and E. Zaccarelli, “Numerical study of neutral and charged microgel suspensions: from single-particle to collective behavior,” arXiv preprint arXiv:2404.04032  (2024).
  • Elancheliyan et al. (2022) R. Elancheliyan, G. Del Monte, E. Chauveau, S. Sennato, E. Zaccarelli,  and D. Truzzolillo, “Role of charge content in the two-step deswelling of poly (n-isopropylacrylamide)-based microgels,” Macromolecules 55, 7526–7539 (2022).
  • Weyer and Denton (2018) T. J. Weyer and A. R. Denton, “Concentration-dependent swelling and structure of ionic microgels: simulation and theory of a coarse-grained model,” Soft Matter 14, 4530–4540 (2018).
  • Cao and Berne (1993) J. Cao and B. J. Berne, “Theory and simulation of polar and nonpolar polarizable fluids,” J. Chem. Phys. 99, 6998–7011 (1993).
  • Calef and Wolynes (1983) D. F. Calef and P. G. Wolynes, “Smoluchowski–vlasov theory of charge solvation dynamics,” J. Chem. Phys. 78, 4145–4153 (1983).
  • Chandra and Bagchi (1990) A. Chandra and B. Bagchi, “Collective orientational relaxation in a dense liquid of ellipsoidal molecules,” Physica A 169, 246–262 (1990).
  • Chandra and Bagchi (1988) A. Chandra and B. Bagchi, “The role of translational diffusion in the polarization relaxation in dense polar liquids,” Chem. Phys. Lett. 151, 47–53 (1988).
  • Motornov et al. (2007) M. Motornov, R. Sheparovych, R. Lupitskyy, E. MacWilliams,  and S. Minko, “Responsive colloidal systems: Reversible aggregation and fabrication of superhydrophobic surfaces,” J. Colloid Interface Sci. 310, 481–488 (2007).
  • Motornov et al. (2011) M. Motornov, H. Royter, R. Lupitskyy, Y. Roiter,  and S. Minko, “Stimuli-responsive hydrogel hollow capsules by material efficient and robust cross-linking-precipitation synthesis revisited,” Langmuir 27, 15305–15311 (2011).
  • Kalaitzidou and Crosby (2008) K. Kalaitzidou and A. J. Crosby, “Adaptive polymer particles,” Appl. Phys. Lett. 93, 041910 (2008).
  • Cheung, Klimov, and Thirumalai (2005) M. S. Cheung, D. Klimov,  and D. Thirumalai, “Molecular crowding enhances native state stability and refolding rates of globular proteins,” Proc. Natl. Acad. Sci. U.S.A. 102, 4753–4758 (2005).
  • Zhou, Rivas, and Minton (2008) H.-X. Zhou, G. Rivas,  and A. P. Minton, “Macromolecular crowding and confinement: Biochemical, biophysical, and potential physiological consequences,” Annu. Rev. Biophys. 37, 375–397 (2008).
  • Hong and Gierasch (2010) J. Hong and L. M. Gierasch, “Macromolecular crowding remodels the energy landscape of a protein by favoring a more compact unfolded state,” J. Am. Chem. Soc. 132, 10445–10452 (2010).
  • Dupuis, Holmstrom, and Nesbitt (2014) N. F. Dupuis, E. D. Holmstrom,  and D. J. Nesbitt, “Molecular-crowding effects on single-molecule rna folding/unfolding thermodynamics and kinetics,” Proc. Natl. Acad. Sci. U.S.A. 111, 8464–8469 (2014).
  • Shin, Cherstvy, and Metzler (2015) J. Shin, A. G. Cherstvy,  and R. Metzler, “Kinetics of polymer loo** with macromolecular crowding: effects of volume fraction and crowder size,” Soft Matter 11, 472–488 (2015).
  • Baul et al. (2021b) U. Baul, N. Götth, M. Bley,  and J. Dzubiella, “Modulating internal transition kinetics in responsive macromolecules by collective crowding,” J. Chem. Phys. 155, 244902 (2021b).
  • Huang et al. (2016) S. Huang, K. Gawlitza, R. von Klitzing, L. Gilson, J. Nowak, S. Odenbach, W. Steffen,  and G. K. Auernhammer, “Microgels at the water/oil interface: In situ observation of structural aging and two-dimensional magnetic bead microrheology,” Langmuir 32, 712–722 (2016).
  • Garbin et al. (2015) V. Garbin, I. Jenkins, T. Sinno, J. C. Crocker,  and K. J. Stebe, “Interactions and stress relaxation in monolayers of soft nanoparticles at fluid-fluid interfaces,” Phys. Rev. Lett. 114, 108301 (2015).
  • Vettorel, Besold, and Kremer (2010) T. Vettorel, G. Besold,  and K. Kremer, “Fluctuating soft-sphere approach to coarse-graining of polymer models,” Soft Matter 6, 2282–2292 (2010).
  • Winkler, Fedosov, and Gompper (2014) R. G. Winkler, D. A. Fedosov,  and G. Gompper, “Dynamical and rheological properties of soft colloid suspensions,” Curr. Opin. Colloid Interface Sci. 19, 594–610 (2014).
  • Moncho-Jordá and Dzubiella (2016) A. Moncho-Jordá and J. Dzubiella, “Swelling of ionic microgel particles in the presence of excluded-volume interactions: a density functional approach,” Phys. Chem. Chem. Phys. 18, 5372–5385 (2016).
  • Karg et al. (2019) M. Karg, A. Pich, T. Hellweg, T. Hoare, L. A. Lyon, J. J. Crassous, D. Suzuki, R. A. Gumerov, S. Schneider, I. I. Potemkin,  and W. Richtering, “Nanogels and microgels: From model colloids to applications, recent developments, and future trends,” Langmuir 35, 6231–6255 (2019).
  • Rovigatti et al. (2019) L. Rovigatti, N. Gnan, A. Ninarello,  and E. Zaccarelli, “Connecting elasticity and effective interactions of neutral microgels: The validity of the hertzian model,” Macromolecules 52, 4895–4906 (2019).
  • Scotti et al. (2019a) A. Scotti, S. Bochenek, M. Brugnoni, M. A. Fernandez-Rodriguez, M. F. Schulte, J. E. Houston, A. P. H. Gelissen, I. I. Potemkin, L. Isa,  and W. Richtering, “Exploring the colloid-to-polymer transition for ultra-low crosslinked microgels from three to two dimensions,” Nat. Commun. 10, 1418 (2019a).
  • Scotti et al. (2019b) A. Scotti, A. R. Denton, M. Brugnoni, J. E. Houston, R. Schweins, I. I. Potemkin,  and W. Richtering, “Deswelling of microgels in crowded suspensions depends on cross-link density and architecture,” Macromolecules 52, 3995–4007 (2019b).
  • Scotti (2021) A. Scotti, “Characterization of the volume fraction of soft deformable microgels by means of small-angle neutron scattering with contrast variation,” Soft Matter 17, 5548–5559 (2021).
  • Murray and Snowden (1995) M. Murray and M. Snowden, “The preparation, characterisation and applications of colloidal microgels,” Adv. Colloid Interface Sci. 54, 73 – 91 (1995).
  • Saunders and Vincent (1999) B. R. Saunders and B. Vincent, “Microgel particles as model colloids: Theory, properties and applications,” Adv. Colloid Interface Sci. 80, 1 – 25 (1999).
  • Fernandez-Nieves et al. (2011) A. Fernandez-Nieves, H. Wyss, J. Mattsson,  and D. A. Weitz, Microgel suspensions: fundamentals and applications (John Wiley & Sons, 2011).
  • Zhou, Tang, and Wu (2015) Y. Zhou, H. Tang,  and P. Wu, “Volume phase transition mechanism of poly [oligo (ethylene glycol) methacrylate] based thermo-responsive microgels with poly (ionic liquid) cross-linkers,” Physical Chemistry Chemical Physics 17, 25525–25535 (2015).
  • Bochenek et al. (2022) S. Bochenek, F. Camerin, E. Zaccarelli, A. Maestro, M. M. Schmidt, W. Richtering,  and A. Scotti, “In-situ study of the impact of temperature and architecture on the interfacial structure of microgels,” Nature Communications 13, 3744 (2022).
  • Nikolov, Fernandez-Nieves, and Alexeev (2020) S. V. Nikolov, A. Fernandez-Nieves,  and A. Alexeev, “Behavior and mechanics of dense microgel suspensions,” Proc. Natl. Acad. Sci. U.S.A. 117, 27096–27103 (2020).
  • Gnan and Zaccarelli (2019) N. Gnan and E. Zaccarelli, “The microscopic role of deformation in the dynamics of soft colloids,” Nat. Phys. 15, 683–688 (2019).
  • Roa et al. (2017) R. Roa, W. K. Kim, M. Kanduč, J. Dzubiella,  and S. Angioletti-Uberti, “Catalyzed bimolecular reactions in responsive nanoreactors,” ACS Catal. 7, 5604–5611 (2017).
  • Roa et al. (2018) R. Roa, S. Angioletti-Uberti, Y. Lu, J. Dzubiella, F. Piazza,  and M. Ballauff, “Catalysis by metallic nanoparticles in solution: Thermosensitive microgels as nanoreactors,” Z. Phys. Chem. 232, 773–803 (2018).
  • Kanduč et al. (2020) M. Kanduč, W. K. Kim, R. Roa,  and J. Dzubiella, “Modeling of stimuli-responsive nanoreactors: rational rate control towards the design of colloidal enzymes,” Mol. Syst. Des. Eng. 5, 602–619 (2020).
  • Hamidi, Azadi, and Rafiei (2008) M. Hamidi, A. Azadi,  and P. Rafiei, “Hydrogel nanoparticles in drug delivery,” Adv. Drug Deliv. Rev. 60, 1638–1649 (2008), 2008 Editors’ Collection.
  • Saunders et al. (2009) B. R. Saunders, N. Laajam, E. Daly, S. Teow, X. Hu,  and R. Stepto, “Microgels: From responsive polymer colloids to biomaterials,” Adv. Colloid Interface Sci. 147-148, 251 – 262 (2009).
  • Vinogradov (2010) S. V. Vinogradov, “Nanogels in the race for drug delivery,” Nanomedicine 5, 165–168 (2010).
  • Moncho-Jordá et al. (2019) A. Moncho-Jordá, A. Germán-Bellod, S. Angioletti-Uberti, I. Adroher-Benítez,  and J. Dzubiella, “Nonequilibrium uptake kinetics of molecular cargo into hollow hydrogels tuned by electrosteric interactions,” ACS Nano 13, 1603–1616 (2019).
  • Moncho-Jordá et al. (2020) A. Moncho-Jordá, A. B. Jódar-Reyes, M. Kanduč, A. Germán-Bellod, J. M. López-Romero, R. Contreras-Cáceres, F. Sarabia, M. García-Castro, H. A. Pérez-Ramírez,  and G. Odriozola, “Scaling laws in the diffusive release of neutral cargo from hollow hydrogel nanoparticles: Paclitaxel-loaded poly(4-vinylpyridine),” ACS Nano 14, 15227–15240 (2020).
  • Pagonabarraga, Cates, and Ackland (2000) I. Pagonabarraga, M. E. Cates,  and G. J. Ackland, “Local size segregation in polydisperse hard sphere fluids,” Phys. Rev. Lett. 84, 911–914 (2000).
  • Buzzacchi, Pagonabarraga, and Wilding (2004) M. Buzzacchi, I. Pagonabarraga,  and N. B. Wilding, “Polydisperse hard spheres at a hard wall,” J. Chem. Phys. 121, 11362–11373 (2004).
  • Pagonabarraga and Cates (2001) I. Pagonabarraga and M. Cates, “A practical density functional for polydisperse polymers,” Europhysics Letters 55, 348 (2001).
  • Moncho-Jordá, Groh, and Dzubiella (2024) A. Moncho-Jordá, S. Groh,  and J. Dzubiella, “External field-driven property localization in liquids of responsive macromolecules,” J. Chem. Phys. 160, 024904 (2024).
  • Moncho-Jordá, Göth, and Dzubiella (2023) A. Moncho-Jordá, N. Göth,  and J. Dzubiella, “Liquid structure of bistable responsive macromolecules using mean-field density-functional theory,” Soft Matter 19, 2832–2846 (2023).
  • te Vrugt, Löwen, and Wittkowski (2020) M. te Vrugt, H. Löwen,  and R. Wittkowski, “Classical dynamical density functional theory: from fundamentals to applications,” Adv. Phys. 69, 121–247 (2020).
  • Marconi and Tarazona (1999) U. M. B. Marconi and P. Tarazona, “Dynamic density functional theory of fluids,” J. Chem. Phys. 110, 8032–8044 (1999).
  • Archer and Evans (2004a) A. J. Archer and R. Evans, “Dynamical density functional theory and its application to spinodal decomposition,” J. Chem. Phys. 121, 4246–4254 (2004a).
  • Royall et al. (2007) C. P. Royall, J. Dzubiella, M. Schmidt,  and A. van Blaaderen, “Nonequilibrium sedimentation of colloids on the particle scale,” Phys. Rev. Lett. 98, 188304 (2007).
  • te Vrugt, Löwen, and Wittkowski (2020) M. te Vrugt, H. Löwen,  and R. Wittkowski, “Classical dynamical density functional theory: from fundamentals to applications,” Adv. Phys. 69, 121–247 (2020).
  • te Vrugt, Bickmann, and Wittkowski (2020) M. te Vrugt, J. Bickmann,  and R. Wittkowski, “Effects of social distancing and isolation on epidemic spreading modeled via dynamical density functional theory,” Nat. Commun. 11, 5576 (2020).
  • Moncho-Jordá and Dzubiella (2020) A. Moncho-Jordá and J. Dzubiella, “Controlling the microstructure and phase behavior of confined soft colloids by active interaction switching,” Phys. Rev. Lett. 125, 078001 (2020).
  • Bley, Dzubiella, and Moncho-Jordá (2021) M. Bley, J. Dzubiella,  and A. Moncho-Jordá, “Active binary switching of soft colloids: stability and structural properties,” Soft Matter 17, 7682–7696 (2021).
  • Bley et al. (2022) M. Bley, P. I. Hurtado, J. Dzubiella,  and A. Moncho-Jordá, “Active interaction switching controls the dynamic heterogeneity of soft colloidal dispersions,” Soft Matter 18, 397–411 (2022).
  • Sato Matsuo and Tanaka (1988) E. Sato Matsuo and T. Tanaka, “Kinetics of discontinuous volume-phase transition of gels,” J. Chem. Phys. 89, 1695–1703 (1988).
  • Suzuki and Suzuki (1995) A. Suzuki and H. Suzuki, “Hysteretic behavior and irreversibility of polymer gels by ph change,” J. Chem. Phys. 103, 4706–4710 (1995).
  • Plaxco et al. (2000) K. W. Plaxco, K. T. Simons, I. Ruczinski,  and D. Baker, “Topology, stability, sequence, and length: Defining the determinants of two-state protein folding kinetics,” Biochemistry 39, 11177–11183 (2000).
  • Murnen et al. (2012) H. K. Murnen, A. R. Khokhlov, P. G. Khalatur, R. A. Segalman,  and R. N. Zuckermann, “Impact of hydrophobic sequence patterning on the coil-to-globule transition of protein-like polymers,” Macromolecules 45, 5229–5236 (2012).
  • Zeng, Ruff, and Pappu (2022) X. Zeng, K. M. Ruff,  and R. V. Pappu, “Competing interactions give rise to two-state behavior and switch-like transitions in charge-rich intrinsically disordered proteins,” Proc. Natl. Acad. Sci. U. S. A. 119, e2200559119 (2022).
  • Choi et al. (2011b) U. B. Choi, J. J. McCann, K. R. Weninger,  and M. E. Bowen, “Beyond the random coil: Stochastic conformational switching in intrinsically disordered proteins,” Structure 19, 566 – 576 (2011b).
  • Dhiman, Jain, and George (2017) S. Dhiman, A. Jain,  and S. J. George, “Transient helicity: Fuel-driven temporal control over conformational switching in a supramolecular polymer,” Angew. Chem. Int. Ed. 56, 1329–1333 (2017).
  • Kang et al. (2015) H. Kang, P. A. Pincus, C. Hyeon,  and D. Thirumalai, “Effects of macromolecular crowding on the collapse of biopolymers,” Phys. Rev. Lett. 114, 068303 (2015).
  • Moncho-Jordá, Göth, and Dzubiella (2023) A. Moncho-Jordá, N. Göth,  and J. Dzubiella, “Liquid structure of bistable responsive macromolecules using mean-field density-functional theory,” Soft Matter 19, 2832–2846 (2023).
  • Likos (2001) C. N. Likos, “Effective interactions in soft condensed matter physics,” Phys. Rep. 348, 267–439 (2001).
  • Bolhuis et al. (2001) P. G. Bolhuis, A. A. Louis, J. P. Hansen,  and E. J. Meijer, “Accurate effective pair potentials for polymer solutions,” J. Chem. Phys. 114, 4296–4311 (2001).
  • Dautenhahn and Hall (1994) J. Dautenhahn and C. K. Hall, “Monte carlo simulation of off-lattice polymer chains: Effective pair potentials in dilute solution,” Macromolecules 27, 5399–5412 (1994).
  • Louis, Bolhuis, and Hansen (2000) A. A. Louis, P. G. Bolhuis,  and J. P. Hansen, “Mean-field fluid behavior of the gaussian core model,” Phys. Rev. E 62, 7961–7972 (2000).
  • Finken, Hansed, and Louis (2003) R. Finken, J. P. Hansed,  and A. A. Louis, “Phase separation of penetrable core mixtures,” J. Stat. Phys. 110, 1015–1037 (2003).
  • Archer and Evans (2001) A. J. Archer and R. Evans, “Binary gaussian core model: Fluid-fluid phase separation and interfacial properties,” Phys. Rev. E 64, 041501 (2001).
  • Archer (2005a) A. J. Archer, “Dynamical density functional theory: binary phase-separating colloidal fluid in a cavity,” J. Phys.-Condes. Matter 17, 1405–1427 (2005a).
  • Archer and Evans (2004b) A. J. Archer and R. Evans, “Dynamical density functional theory and its application to spinodal decomposition,” J. Chem. Phys. 121, 4246–4254 (2004b).
  • Archer, Likos, and Evans (2004) A. J. Archer, C. N. Likos,  and R. Evans, “Soft-core binary fluid exhibiting a λ𝜆\lambdaitalic_λ-line and freezing to a highly delocalized crystal,” J. Phys.-Condes. Matter 16, L297 (2004).
  • Archer (2005b) A. J. Archer, “Dynamical density functional theory: phase separation in a cavity and the influence of symmetry,” J. Phys.-Condes. Matter 17, S3253–S3258 (2005b).
  • Schmidt (2022) M. Schmidt, “Power functional theory for many-body dynamics,” Rev. Mod. Phys. 94, 015007 (2022).
  • Milster et al. (2023) S. Milster, A. Darwish, N. Göth,  and J. Dzubiella, “Synergistic chemomechanical dynamics of feedback-controlled microreactors,” Phys. Rev. E 108, L042601 (2023).