Collapse/expansion dynamics and actuation of pH-responsive nanogels

Jiaxing Yuan Research Center for Advanced Science and Technology, University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8904, Japan    Tine Curk [email protected] Department of Materials Science & Engineering, Johns Hopkins University, Baltimore, Maryland 21218, USA
(May 29, 2024)
Abstract

Polyelectrolyte (PE) hydrogels can dynamically respond to external stimuli, such as changes in pH and temperature, which benefits their use for smart materials and nanodevices with tunable properties. We investigate equilibrium conformations and phase transition dynamics of pH-responsive nanogels using hybrid molecular dynamics/Monte Carlo simulations with full consideration of electrostatic and hydrodynamic interactions. We demonstrate that PE nanogels exhibit a closed-loop phase behavior with a discontinuous swelling–collapse transition that occurs only at intermediate pH values. A 50 nm nanogel particle close to a critical point functions as a pH-driven actuator with a microsecond conformational response and work density 105J/m3absentsuperscript105superscriptJ/m3\approx 10^{5}~{}\textnormal{J/m}^{3}≈ 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT J/m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, an order of magnitude larger than skeletal muscles. The collapse/expansion time scales as L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the power density scales as L2superscript𝐿2L^{-2}italic_L start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT where L𝐿Litalic_L is the linear size of the gel. Our work provides fundamental insight into phase behavior and non-equilibrium dynamics of the swelling–collapse transition, and our method enables the investigation of charge–structure–hydrodynamic coupling in soft materials.

Charge regulation, electrostatic interaction, polyelectrolyte hydrogel, discontinuous transition

Polyelectrolyte (PE) hydrogels are three-dimensional networks of cross-linked charged polymers that have found widespread industrial applications Karg et al. (2019). The conformational response of PE gels to external stimuli, such as variation in temperature van der Linden and Westerweel (2013); Ge et al. (2021), pH Gupta et al. (2002); Glazer et al. (2012); Han et al. (2020), salt concentration Shao et al. (2021), and electrical or magnetic fields Schreyer et al. (2000), makes them prime building blocks for the design of smart materials and nanodevices such as drug delivery vehicles Gupta et al. (2002), and actuators for soft robotics Morales et al. (2014); Li et al. (2020). Through the incorporation of stimuli-responsive components such as ionizable groups, hydrogels undergo shape changes and perform mechanical work in response to triggers such as changes in temperature and pH Li et al. (2022). Understanding both the thermodynamics and kinetics of the swelling–collapse transition of hydrogels is crucial for improving our ability to precisely control their shape change and actuation.

A polymer gel typically undergoes a swelling–collapse transition upon a change in temperature. This transition is continuous for a free-standing uncharged gel Quesada-Pérez et al. (2011), but can become discontinuous for gels under tension Dus̆ek and Patterson (1968) or highly charged PE gels in low-dielectric solvents Tanaka et al. (1980); Yan and de Pablo (2003). Moreover, hydrogels are typically weak acids/bases where the charge of the hydrogel is not constant but depends on the gel conformation due to the charge regulation (CR) effect Curk and Luijten (2021); Curk et al. (2022). Charge regulation leads to a discontinuous swelling–collapse transition of pH-responsive hydrogels, which terminates in a critical point at a finite pH value, as is predicted by mean-field theory Tanaka et al. (1980); Polotsky et al. (2013); Zheng et al. (2023). While extensive research has been conducted on the phase diagram of bulk PE hydrogels Tanaka et al. (1980); Polotsky et al. (2013); Yan and de Pablo (2003); Zheng et al. (2023), there is a notable lack of attention given to the swelling–collapse kinetics. This is especially the case for nanogels since conventional microscopic techniques struggle to capture the time evolution of nanoscale conformations Keidel et al. (2018); Dallari et al. (2024). In particular, nanogels possess substantial surface-to-volume ratio Zhou et al. (2020); Wang et al. (2021) and offer faster dynamical response to the external stimuli Asadian-Birjand et al. (2012), making them highly promising for the emerging biomedical and nanotechnology applications.

In this Letter, we explore the charge–structure–hydrodynamic coupling in a PE nanogel. We use hybrid molecular dynamics (MD)/MC simulations to investigate equilibrium conformations and transition dynamics of a pH-responsive nanogel, focusing on the impact of electrostatic and hydrodynamic interactions on the collapse and expansion kinetics. From these we derive design guidelines for engineering pH-driven nanoactuators.

To understand the actuation behavior, we first calculate the equilibrium phase diagram of a nanogel particle. We employ a coarse-grained PE model where all charges and all ionizable sites in the hydrogel are explicitly represented, allowing accurate calculation of electrostatic interactions. Moreover, the model captures the dynamic interplay between protonation, gel conformation, electrostatic interaction, counter-ion osmotic pressure, and excluded-volume effects. The anionic gel is immersed in an aqueous solution of Bjerrum length lB=βq02/(4πεsol)subscript𝑙B𝛽superscriptsubscript𝑞024𝜋subscript𝜀soll_{\textnormal{B}}=\beta q_{0}^{2}/(4\pi\varepsilon_{\textnormal{sol}})italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT = italic_β italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 4 italic_π italic_ε start_POSTSUBSCRIPT sol end_POSTSUBSCRIPT ), where β=1/(kBT)𝛽1subscript𝑘B𝑇\beta=1/(k_{\textnormal{B}}T)italic_β = 1 / ( italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T ), kBsubscript𝑘Bk_{\textnormal{B}}italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT is Boltzmann constant, T𝑇Titalic_T is the absolute temperature, q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denotes the elementary charge, and εsolsubscript𝜀sol\varepsilon_{\text{sol}}italic_ε start_POSTSUBSCRIPT sol end_POSTSUBSCRIPT is the solvent permittivity, resulting in lB=0.72nmsubscript𝑙B0.72nml_{\textnormal{B}}=0.72\,\textnormal{nm}italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT = 0.72 nm for a hydrogel at room temperature. The gel consists of ionizable bead-spring PE chains cross-linked into a gel network with crosslinking distance Npsubscript𝑁pN_{\textnormal{p}}italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT [Fig. 1(a)–(b)].

Ions and monomers are represented as spheres of diameter σ=lB𝜎subscript𝑙B\sigma=l_{\textnormal{B}}italic_σ = italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT, which is a typical size of hydrated ions or ionizable groups. The short-range attraction between PE monomers is represented by the Lennard-Jones potential with strength εmmsubscript𝜀mm\varepsilon_{\textnormal{mm}}italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT, which determines the solvent quality, and cutoff distance rcut=3σsubscript𝑟cut3𝜎r_{\text{cut}}=3\sigmaitalic_r start_POSTSUBSCRIPT cut end_POSTSUBSCRIPT = 3 italic_σ. Each ionizable site (monomer) is subject to the acid dissociation reaction, A0A+H+superscriptA0superscriptAsuperscriptH\textnormal{A}^{0}\rightleftharpoons\textnormal{A}^{-}+\textnormal{H}^{+}A start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ⇌ A start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT + H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, where A0superscriptA0\textnormal{A}^{0}A start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and AsuperscriptA\textnormal{A}^{-}A start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT are the neutral and dissociated states with H+superscriptH\textnormal{H}^{+}H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT the dissociated proton. The probability αisubscript𝛼𝑖\alpha_{i}italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT that a site i𝑖iitalic_i at position 𝐫isubscript𝐫𝑖\mathbf{r}_{i}bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is charged (dissociated) is determined by Podgornik (2018); Curk and Luijten (2021),

αi1αi=10pHpKaeβψ(𝐫i)q0,subscript𝛼𝑖1subscript𝛼𝑖superscript10pHpsubscript𝐾asuperscript𝑒𝛽𝜓subscript𝐫𝑖subscript𝑞0\frac{\alpha_{i}}{1-\alpha_{i}}=10^{\textnormal{pH}-\textnormal{p}K_{% \textnormal{a}}}e^{-\beta\psi(\mathbf{r}_{i})q_{0}}\;,divide start_ARG italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_α start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG = 10 start_POSTSUPERSCRIPT pH - p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - italic_β italic_ψ ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT , (1)

with pKapsubscript𝐾a\textnormal{p}K_{\textnormal{a}}p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT the equilibrium dissociation constant and ψ(𝐫i)𝜓subscript𝐫𝑖\psi(\mathbf{r}_{i})italic_ψ ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) the electrostatic potential at site i𝑖iitalic_i. We model all charged entities explicitly and perform grand-canonical MC exchange of dissociated ions (H+superscriptH\textnormal{H}^{+}H start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT) and monovalent salt cations (S+superscriptS\textnormal{S}^{+}S start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT) and anions (SsuperscriptS\textnormal{S}^{-}S start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT) with a reservoir at a given pH and monovalent salt concentration cs=103subscript𝑐ssuperscript103c_{\textnormal{s}}=10^{-3}italic_c start_POSTSUBSCRIPT s end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPTM. Electrostatic interactions are calculated using particle–particle particle–mesh algorithm Hockney and Eastwood (1988) with relative force accuracy 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT.

The system configurations are evolved employing the standard velocity-Verlet MD algorithm, while ionization states [Eq. (1)] and ion exchange are sampled using an efficient charge regulation MC (CR-MC) solver Curk et al. (2022). To investigate the equilibrium phase diagram, we employ the standard Langevin dynamics simulations. The phase transition dynamics of PE are influenced by hydrodynamic interactions (HI) Kikuchi et al. (2005); Kamata et al. (2009); Yuan et al. (2022); Yuan and Tanaka (2024). Thus, for the study of non-equilibrium collapse/expansion dynamics of nanogels, we account for HI by combining the CR-MC solver with a method based on dissipative particle dynamics (DPD) Groot and Warren (1997) that couples the solute (gel) particles to the DPD solvent (DPDS) Curk (2024). Additional simulation details are provided in Supplemental Material Sup .

Refer to caption
Figure 1: Equilibrium phase diagram of a nanogel. (a,b) Typical expanded (ΔpH=6ΔpH6\Delta\textnormal{pH}=6roman_Δ pH = 6) and collapsed (ΔpH=1ΔpH1\Delta\textnormal{pH}=1roman_Δ pH = 1) conformations at εmm=kBTsubscript𝜀mmsubscript𝑘B𝑇\varepsilon_{\textnormal{mm}}=k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T showing negatively charged dissociated (blue) and neutral (white) acid groups, free cations (red), and anions (metallic blue), note ΔpH=pHpKaΔpHpHpsubscript𝐾a\Delta\textnormal{pH}=\textnormal{pH}-\textnormal{p}K_{\textnormal{a}}roman_Δ pH = pH - p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT. (c,d) Average gel volume Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT and degree of ionization α𝛼\alphaitalic_α showing a sharp transition at intermediate ΔpHΔpH\Delta\textnormal{pH}roman_Δ pH and εmmsubscript𝜀mm\varepsilon_{\textnormal{mm}}italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT. Triangles in (c) and (d) mark the locations of the peaks of equilibrium fluctuations in Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT and α𝛼\alphaitalic_α. The dashed black lines mark the discontinuous transition and the arrows represent the kinetic pathway explored below. (e) Average α𝛼\alphaitalic_α displaying a closed-loop phase diagram with a central discontinuous transition. (f) The transition free-energy barrier along the dashed black line in (c,d). The volume Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT is obtained from the convex hull of the gel monomers. Parameters: width Nx=Ny=2subscript𝑁𝑥subscript𝑁𝑦2N_{x}=N_{y}=2italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 2 and length Nz=6subscript𝑁𝑧6N_{z}=6italic_N start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 6 unit cells of size Np=10subscript𝑁p10N_{\textnormal{p}}=10italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 10, containing a total of 1395 ionizable monomers.

We initially focus on an elongated nanogel which is a prototypical system for hydrogel-based soft actuators and artificial muscles Park and Kim (2020). Based on experimental observations Tanaka et al. (1980) and theoretical Polotsky et al. (2013); Zheng et al. (2023) predictions, we anticipate that a hydrogel exhibits a critical point beyond which the collapse–swelling transition is discontinuous. We compute the pH–εmm/kBTsubscript𝜀mmsubscript𝑘B𝑇\varepsilon_{\textnormal{mm}}/k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T phase diagram of a nanogel and find a sharp collapse–swelling transition at intermediate pH and εmm/kBTsubscript𝜀mmsubscript𝑘B𝑇\varepsilon_{\textnormal{mm}}/k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T values [Fig. 1(c,d)]. The discontinuous transition arises due to CR coupling between the gel conformation and ionization (Eq. (1)), leading to coexistence between an expanded charged gel and a collapsed uncharged gel. However, this CR coupling and the discontinuous transition disappear in the limits pH±pHplus-or-minus\textnormal{pH}\to\pm\inftypH → ± ∞ at which the charge on the gel is constant (α1𝛼1\alpha\to 1italic_α → 1 or α0𝛼0\alpha\to 0italic_α → 0). Fluctuation analysis of Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT and α𝛼\alphaitalic_α (Fig. S1 in Sup ) further shows two distinct peaks that indicate tentative locations of the two critical points [marked by triangles in Fig. 1(c)–(d)]. The peak locations are consistent for fluctuations in gel volume Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT and ionization α𝛼\alphaitalic_α. Thus, interestingly, the discontinuous region is limited to intermediate pH values, implying a closed-loop phase diagram [Fig. 1(e)]. Free energy calculations [Fig. 1(f); also see Fig. S2 in Sup ] confirm two stable states separated by a barrier at intermediate pH. Finite-size scaling shows the barrier increases with increasing gel volume (Fig. S3 in Sup ), which further supports that the conformational change is a discontinuous transition at intermediate pH values. On the other hand, the barrier disappears at both high and low pH values indicating that a continuous transition occurs for both neutral (high pH, α0similar-to𝛼0\alpha\sim 0italic_α ∼ 0) and fully ionized (low pH, α1similar-to𝛼1\alpha\sim 1italic_α ∼ 1) nanogel (Fig. S4 in Sup ).

The location of the transition is not significantly affected by varying the shape of the gel or the crosslink distance (Fig. S5 in Sup ), which implies that the phase diagram in Fig. 1(c)–(d) is robust and is expected to apply to randomly cross-linked PE hydrogels. Indeed, our findings can explain why the experimentally observed hydrogel volume transition is sharpest at intermediate temperature and pH Kang and Bae (2001).

Having calculated the equilibrium phase diagram of the nanogel, we turn to kinetics and design a pH-responsive actuator. We introduce a constant force with magnitude fesubscript𝑓ef_{\textnormal{e}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT to the two ends of the hydrogel [Fig. 2(a)] and apply a zigzag time-varying pH cycle between pH0=pKasubscriptpH0psubscript𝐾a\textnormal{pH}_{\textnormal{0}}=\textnormal{p}K_{\textnormal{a}}pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT and pH1=pKa+6subscriptpH1psubscript𝐾a6\textnormal{pH}_{1}=\textnormal{p}K_{\textnormal{a}}+6pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT + 6 with ramp rate ω=d(pH)/dt𝜔𝑑pH𝑑𝑡\omega={d(\textnormal{pH})}/{dt}italic_ω = italic_d ( pH ) / italic_d italic_t (see Sup for details). The total time to complete a full cycle is tf=2(pH1pH0)/ωsubscript𝑡f2subscriptpH1subscriptpH0𝜔t_{\textnormal{f}}=2(\textnormal{pH}_{\textnormal{1}}-\textnormal{pH}_{% \textnormal{0}})/\omegaitalic_t start_POSTSUBSCRIPT f end_POSTSUBSCRIPT = 2 ( pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_ω. Guided by the phase diagram [Fig. 1(c)–(d)] we chose εmm=1kBTsubscript𝜀mm1subscript𝑘B𝑇\varepsilon_{\textnormal{mm}}=1k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT = 1 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T, which is close to the peak in equilibrium fluctuations and is thus expected to result in a sharp and fast response without exhibiting strong hysteresis. Hydrodynamics is resolved using the DPDS method Curk (2024) that models an aqueous solution with time unit τ=0.029ns𝜏0.029ns\tau=0.029~{}\textnormal{ns}italic_τ = 0.029 ns, while the ionization kinetics is modeled using the CR-MC method Curk et al. (2022) with realistic ionization rates kd3×106s1subscript𝑘d3superscript106superscripts1k_{\textnormal{d}}\approx 3\times 10^{6}~{}\textnormal{s}^{-1}italic_k start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ≈ 3 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT Kanzaki et al. (2014).

Refer to caption
Figure 2: Polyelectrolyte hydrogel as a pH-responsive soft actuator. (a) An external force fesubscript𝑓ef_{\textnormal{e}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT is applied to the ends of the gel. (b,c) The response of a weak polyelectrolyte gel (pKa=4psubscript𝐾a4\textnormal{p}K_{\textnormal{a}}=4p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT = 4) under pH ram** rate ω𝜔\omegaitalic_ω at fe=0subscript𝑓e0f_{\textnormal{e}}=0italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = 0 (b) and fe=9kBT/lBsubscript𝑓e9subscript𝑘B𝑇subscript𝑙Bf_{\textnormal{e}}=9k_{\textnormal{B}}T/l_{\textnormal{B}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = 9 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T / italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT (c). (d) The corresponding power output, P=fedRee/dt𝑃subscript𝑓e𝑑subscript𝑅ee𝑑𝑡P=-f_{\textnormal{e}}dR_{\textnormal{ee}}/dtitalic_P = - italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT italic_d italic_R start_POSTSUBSCRIPT ee end_POSTSUBSCRIPT / italic_d italic_t, at fe=9kBT/lBsubscript𝑓e9subscript𝑘B𝑇subscript𝑙Bf_{\textnormal{e}}=9k_{\textnormal{B}}T/l_{\textnormal{B}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT = 9 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T / italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT. (e) The work performed against the external force upon changing pH from pH1=pKa+6subscriptpH1psubscript𝐾a6\textnormal{pH}_{1}=\textnormal{p}K_{\textnormal{a}}+6pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT + 6 to pH0=pKasubscriptpH0psubscript𝐾a\textnormal{pH}_{0}=\textnormal{p}K_{\textnormal{a}}pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = p italic_K start_POSTSUBSCRIPT a end_POSTSUBSCRIPT under w0𝑤0w\to 0italic_w → 0. (f) The maximum power Pmaxsubscript𝑃maxP_{\text{max}}italic_P start_POSTSUBSCRIPT max end_POSTSUBSCRIPT as a function of external force at the ram** rate ω=0.0006τ1𝜔0.0006superscript𝜏1\omega=0.0006\tau^{-1}italic_ω = 0.0006 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and ω=0.0002τ1𝜔0.0002superscript𝜏1\omega=0.0002\tau^{-1}italic_ω = 0.0002 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The error bars show standard errors.

We find that the end-to-end distance Reesubscript𝑅eeR_{\textnormal{ee}}italic_R start_POSTSUBSCRIPT ee end_POSTSUBSCRIPT of the gel follows the pH cycle and shows hysteresis that becomes more pronounced at faster cycling rates [Fig. 2(b)]. The response under external force [Fig. 2(c)] remains qualitatively the same, but conformations are more elongated due to the stretching by the external force. The maximum instantaneous power is Pmax0.1kBT/τsubscript𝑃max0.1subscript𝑘B𝑇𝜏P_{\textnormal{max}}\approx 0.1k_{\textnormal{B}}T/\tauitalic_P start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≈ 0.1 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T / italic_τ at the high-frequency limit (ω=0.0006τ1𝜔0.0006superscript𝜏1\omega=0.0006\tau^{-1}italic_ω = 0.0006 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) allowable by protonation dynamics Kanzaki et al. (2014) [Fig. 2(d)]. The response time of the gel at lower rate ω=0.0002τ1𝜔0.0002superscript𝜏1\omega=0.0002\tau^{-1}italic_ω = 0.0002 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is τr3/ω0.4μssubscript𝜏r3𝜔0.4𝜇s\tau_{\textnormal{r}}\approx 3/\omega\approx 0.4~{}\mu\textnormal{s}italic_τ start_POSTSUBSCRIPT r end_POSTSUBSCRIPT ≈ 3 / italic_ω ≈ 0.4 italic_μ s, using the DPD timescale τ0.029𝜏0.029\tau\approx 0.029italic_τ ≈ 0.029 ns Curk (2024), which is at least 4 orders of magnitude faster than typical values for mammalian skeletal muscles  Ranatunga (1998); Mirfakhrai et al. (2007) or macroscopic soft actuators Kim et al. (2015); Ma et al. (2020a). This fast response is a consequence of the nanoscale width of the gel which permits fast transport of solvent.

The work W𝑊Witalic_W performed during the contraction of the gel is W=feΔRee𝑊subscript𝑓eΔsubscript𝑅eeW=f_{\textnormal{e}}\Delta R_{\mathrm{ee}}italic_W = italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT roman_Δ italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPT, with the change in the end to end distance ΔRee=[Ree(pH1)Ree(pH0)]Δsubscript𝑅eedelimited-[]subscript𝑅eesubscriptpH1subscript𝑅eesubscriptpH0\Delta R_{\mathrm{ee}}=[R_{\textnormal{ee}}(\textnormal{pH}_{1})-R_{% \textnormal{ee}}(\textnormal{pH}_{0})]roman_Δ italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPT = [ italic_R start_POSTSUBSCRIPT ee end_POSTSUBSCRIPT ( pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - italic_R start_POSTSUBSCRIPT ee end_POSTSUBSCRIPT ( pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ]. W𝑊Witalic_W reaches a maximum at force fe18kBT/lBsuperscriptsubscript𝑓e18subscript𝑘B𝑇subscript𝑙Bf_{\textnormal{e}}^{*}\approx 18k_{\textnormal{B}}T/l_{\textnormal{B}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≈ 18 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T / italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT and decreases at larger forces that prevent the gel from fully collapsing [Fig. 2(e)]. The optimal load translates to stress σf=fe/A5×105Pasubscript𝜎𝑓superscriptsubscript𝑓e𝐴5superscript105Pa\sigma_{f}=f_{\textnormal{e}}^{*}/A\approx 5\times 10^{5}~{}\textnormal{Pa}italic_σ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_A ≈ 5 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT Pa, with the cross-section area, ANp2NxNyσ2=400lB2𝐴superscriptsubscript𝑁p2subscript𝑁𝑥subscript𝑁𝑦superscript𝜎2400superscriptsubscript𝑙B2A\approx N_{\textnormal{p}}^{2}N_{x}N_{y}\sigma^{2}=400l_{\textnormal{B}}^{2}italic_A ≈ italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 400 italic_l start_POSTSUBSCRIPT B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, under which conditions the gel can elongate twofold. Moreover, the maximum work Wsuperscript𝑊W^{*}italic_W start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT at the force fesuperscriptsubscript𝑓ef_{\textnormal{e}}^{*}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT is W400kBTsuperscript𝑊400subscript𝑘B𝑇W^{*}\approx 400k_{\textnormal{B}}Titalic_W start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≈ 400 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T [Fig. 2(e)], corresponding to a work density of W/Vgel150kJ/m3superscript𝑊subscript𝑉gel150kJsuperscriptm3W^{*}/V_{\textnormal{gel}}\approx 150\textnormal{kJ}/\textnormal{m}^{3}italic_W start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT / italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT ≈ 150 kJ / m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. Thus, the maximum strain of the pH-driven gel actuator is comparable to that of skeletal muscles, while the stress and work density are an order of magnitude larger Mirfakhrai et al. (2007). The maximum instantaneous power Pmaxsubscript𝑃maxP_{\textnormal{max}}italic_P start_POSTSUBSCRIPT max end_POSTSUBSCRIPT at the high frequency limit (ω=0.0006τ1𝜔0.0006superscript𝜏1\omega=0.0006\tau^{-1}italic_ω = 0.0006 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) also exhibits a peak close to fesuperscriptsubscript𝑓ef_{\textnormal{e}}^{*}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT [Fig. 2(f)], suggesting that the nanogel actuator has the potential to yield favorable performance in terms of both work and power even at fast cycling rates.

To explore the optimal design of the actuator we calculate the work, power, and hysteresis over a full spectrum of εmm/kBTsubscript𝜀mmsubscript𝑘B𝑇\varepsilon_{\textnormal{mm}}/k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T and fesubscript𝑓ef_{\textnormal{e}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT (Fig. 3). The equilibrium work performed per contraction, W(ω0)𝑊𝜔0W(\omega\to 0)italic_W ( italic_ω → 0 ), is largest in the discontinuous region of the phase diagram and strong forces (under external force, the discontinuous region moves to larger εmmsubscript𝜀mm\varepsilon_{\textnormal{mm}}italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT) [Fig. 3(a)]. At fast cycling rates, however, the response is limited by the viscous drag and we find the mean power per contraction,

P¯(ω)=W(ω)ωpH1pH0,¯𝑃𝜔𝑊𝜔𝜔subscriptpH1subscriptpH0\bar{P}(\omega)=W(\omega)\frac{\omega}{\textnormal{pH}_{1}-\textnormal{pH}_{0}% }\;,over¯ start_ARG italic_P end_ARG ( italic_ω ) = italic_W ( italic_ω ) divide start_ARG italic_ω end_ARG start_ARG pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , (2)

is maximal at intermediate fesubscript𝑓ef_{\textnormal{e}}italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT and εmmsubscript𝜀mm\varepsilon_{\textnormal{mm}}italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT [Fig. 3(b)]. This observation underscores the crucial role of hydrodynamic effects: the slow relaxation dynamics at εmm/kBT>2subscript𝜀mmsubscript𝑘B𝑇2\varepsilon_{\text{mm}}/k_{\text{B}}T>2italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T > 2 fail to keep pace with the pH ram** rate. Note that the power plot is obtained at the high frequency limit allowable by protonation dynamics, ω=0.0006/τ𝜔0.0006𝜏\omega=0.0006/\tauitalic_ω = 0.0006 / italic_τ Kanzaki et al. (2014), thus the work and power plots in Figs. 3(a)–3(b) effectively show the low and high frequency limits, respectively.

To quantify the hysteresis we calculate the effective width H𝐻Hitalic_H between the forward and reverse process,

H(ω)=1ΔRee(ω)Reed(pH)𝐻𝜔1Δsubscript𝑅ee𝜔contour-integralsubscript𝑅ee𝑑pHH(\omega)=\frac{1}{\Delta R_{\textnormal{ee}}(\omega)}\oint R_{\mathrm{ee}}d(% \textnormal{pH})italic_H ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG roman_Δ italic_R start_POSTSUBSCRIPT ee end_POSTSUBSCRIPT ( italic_ω ) end_ARG ∮ italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPT italic_d ( pH ) (3)

as the integration area of the closed loop in Reesubscript𝑅eeR_{\mathrm{ee}}italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPTpH divided by the total change in size ΔRee(ω)Δsubscript𝑅ee𝜔\Delta R_{\mathrm{ee}}(\omega)roman_Δ italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPT ( italic_ω ) [Fig. 3(c)]. Hysteresis is lowest at εmm/kBT1subscript𝜀mmsubscript𝑘B𝑇1\varepsilon_{\textnormal{mm}}/k_{\textnormal{B}}T\leq 1italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T ≤ 1, which coincides with the continuous region of the phase diagram [cf. Fig. 1(c)]. The hysteresis effect becomes significantly stronger under a faster ram** rate of ω=0.003τ1𝜔0.003superscript𝜏1\omega=0.003\tau^{-1}italic_ω = 0.003 italic_τ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [Fig. S6(c) in Sup ] at which ionization equilibrium cannot be achieved (kd(pH1pH0)/ω<1subscript𝑘dsubscriptpH1subscriptpH0𝜔1k_{\textnormal{d}}(\textnormal{pH}_{\textnormal{1}}-\textnormal{pH}_{% \textnormal{0}})/\omega<1italic_k start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ( pH start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - pH start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_ω < 1).

These findings indicate that the overall optimal condition for actuator performance is close to the critical point, εmm/kBT=1subscript𝜀mmsubscript𝑘B𝑇1\varepsilon_{\text{mm}}/k_{\text{B}}T=1italic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T = 1, resulting in fast shape changes with appreciable power output, but without strong hysteresis.

Refer to caption
Figure 3: Performance of polyelectrolyte nanogel actuator. (a) Equilibrium work per contraction W=feΔRee(ω0)𝑊subscript𝑓eΔsubscript𝑅ee𝜔0W=f_{\textnormal{e}}\Delta R_{\mathrm{ee}}(\omega\to 0)italic_W = italic_f start_POSTSUBSCRIPT e end_POSTSUBSCRIPT roman_Δ italic_R start_POSTSUBSCRIPT roman_ee end_POSTSUBSCRIPT ( italic_ω → 0 ), (b) mean power P¯¯𝑃\bar{P}over¯ start_ARG italic_P end_ARG, and (c) hysteresis H𝐻Hitalic_H (see Eq. 3) at ram** rate ω=0.0006/τ𝜔0.0006𝜏\omega=0.0006/\tauitalic_ω = 0.0006 / italic_τ. (d) The collapse timescale tcolsubscript𝑡colt_{\text{col}}italic_t start_POSTSUBSCRIPT col end_POSTSUBSCRIPT and the expansion timescale texpsubscript𝑡expt_{\text{exp}}italic_t start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT of a neutral and fully charged gel as a function of gel size L𝐿Litalic_L. The dashed lines have a slope of 2.

As the timescales of collapse tcolsubscript𝑡colt_{\textnormal{col}}italic_t start_POSTSUBSCRIPT col end_POSTSUBSCRIPT and expansion texpsubscript𝑡expt_{\textnormal{exp}}italic_t start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT directly determine the maximum power, it is also essential to understand the impact of nanogel size L𝐿Litalic_L on the swelling–collapse dynamics. We consider a cubic nanogel where Nx=Ny=Nz=Nsubscript𝑁𝑥subscript𝑁𝑦subscript𝑁𝑧𝑁N_{x}=N_{y}=N_{z}=Nitalic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_N at crosslinking distance Np=10subscript𝑁p10N_{\textnormal{p}}=10italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 10 and calculate the expansion/collapse timescales as a function of size L=NNp𝐿𝑁subscript𝑁pL=NN_{\textnormal{p}}italic_L = italic_N italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT. The red and blue arrows in Fig. 1(c) denote the pathway of four different cases: the collapse or expansion of fully charged and neutral gels (see Sup for the detailed protocols to induce collapse/expansion dynamics).

Interestingly, we find that both the collapse and expansion timescales exhibit scaling with the gel size L𝐿Litalic_L as tcolL2similar-tosubscript𝑡colsuperscript𝐿2t_{\text{col}}\sim L^{2}italic_t start_POSTSUBSCRIPT col end_POSTSUBSCRIPT ∼ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and texpL2similar-tosubscript𝑡expsuperscript𝐿2t_{\text{exp}}\sim L^{2}italic_t start_POSTSUBSCRIPT exp end_POSTSUBSCRIPT ∼ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [Fig. 3(d)] for both charged and neutral gels (see Fig. S7 and Sup for details on the determination of timescales). This aligns with a theoretical prediction for a swelling timescale of a neutral gel Tanaka and Fillmore (1979). Moreover, the expansion of a neutral gel is much slower than its collapse [Fig. 3(d)], which is consistent with experimental data Dallari et al. (2024). Given the work WL3similar-to𝑊superscript𝐿3W\sim L^{3}italic_W ∼ italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (the force is proportional to L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and the change in length to L𝐿Litalic_L), it follows that the power density of the actuator scales as P/VgelL2Vgel2/3similar-to𝑃subscript𝑉gelsuperscript𝐿2similar-tosuperscriptsubscript𝑉gel23P/V_{\textnormal{gel}}\sim L^{-2}\sim V_{\textnormal{gel}}^{-2/3}italic_P / italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT ∼ italic_L start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ∼ italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT. For the simulated nanogel actuator [Fig. 2(f)], the peak power is 0.5kBT/τ8×10110.5subscript𝑘B𝑇𝜏8superscript10110.5k_{\textnormal{B}}T/\tau\approx 8\times 10^{-11}0.5 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T / italic_τ ≈ 8 × 10 start_POSTSUPERSCRIPT - 11 end_POSTSUPERSCRIPTW and the gel volume is VgelNxNyNzNp3σ3105μm3subscript𝑉gelsubscript𝑁𝑥subscript𝑁𝑦subscript𝑁𝑧superscriptsubscript𝑁p3superscript𝜎3superscript105𝜇superscriptm3V_{\text{gel}}\approx N_{x}N_{y}N_{z}N_{\textnormal{p}}^{3}\sigma^{3}\approx 1% 0^{-5}\mu\text{m}^{3}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT ≈ italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_μ m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, thus the maximal power density is Pmax/Vgel105W/μm3subscript𝑃maxsubscript𝑉gelsuperscript105𝑊𝜇superscriptm3P_{\textnormal{max}}/V_{\textnormal{gel}}\approx 10^{-5}W/\mu\text{m}^{3}italic_P start_POSTSUBSCRIPT max end_POSTSUBSCRIPT / italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT italic_W / italic_μ m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. Using the scaling we predict that for a microgel particle of Vgel=103μm3subscript𝑉gelsuperscript103𝜇superscriptm3V_{\text{gel}}=10^{3}\mu\text{m}^{3}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_μ m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, the power density is about 0.04W/mm30.04superscriptW/mm30.04\text{W/mm}^{3}0.04 W/mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. This showcases remarkable performance compared to the current hydrogel actuators Park et al. (2024). For instance, a bioinspired strong contractile hydrogel with a similar volume of 103μm3similar-toabsentsuperscript103𝜇superscriptm3\sim 10^{3}\mu\text{m}^{3}∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_μ m start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT demonstrates a power density of 5×108W/mm3similar-toabsent5superscript108superscriptW/mm3\sim 5\times 10^{-8}\text{W/mm}^{3}∼ 5 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT W/mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT Ma et al. (2020b).

We evaluate the effects of ion diffusion and protonation kinetics on the actuator performance in Appendix A. Our model does not include effects related to orientational ordering of water and related changes of permittivity close to ions, we discuss these aspects in Appendix B.

In summary, by dynamically modeling the charge–structure coupling via charge regulation of individual dissociable groups, we have calculated the pH–εmm/Tsubscript𝜀mm𝑇\varepsilon_{\textnormal{mm}}/Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_T and Vgelsubscript𝑉gelV_{\textnormal{gel}}italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPTεmm/Tsubscript𝜀mm𝑇\varepsilon_{\textnormal{mm}}/Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_T phase diagrams of a nanogel particle. We demonstrate a discontinuous swelling–collapse transition for intermediate values of pH and interaction strength εmm/Tsubscript𝜀mm𝑇\varepsilon_{\textnormal{mm}}/Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT / italic_T. The phase diagrams serve as a guide for predicting and engineering nanogel materials. By modeling the dynamics of collapse/expansion, including full hydrodynamic interactions, we demonstrate that the charge–structure–hydrodynamic coupling enables the gel to serve as an effective pH-responsive nanoactuator; the maximum strain of the proton-driven gel actuator is comparable to that of skeletal muscles, while the tensile stress and work density are an order of magnitude larger.

Crucially, we show that the optimal performance in terms of high power and low hysteresis is found near a critical point, εmmkBTsubscript𝜀mmsubscript𝑘B𝑇\varepsilon_{\textnormal{mm}}\approx k_{\textnormal{B}}Titalic_ε start_POSTSUBSCRIPT mm end_POSTSUBSCRIPT ≈ italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T, while the power density decreases with the gel volume as P/VgelVgel2/3similar-to𝑃subscript𝑉gelsuperscriptsubscript𝑉gel23P/V_{\textnormal{gel}}\sim V_{\textnormal{gel}}^{-2/3}italic_P / italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT ∼ italic_V start_POSTSUBSCRIPT gel end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT. These scaling results show that the response of bulk hydrogels is very slow due to solvent diffusion. We hypothesize that a fast response in a bulk material could be achieved by introducing holes in a bulk gel through which solvent could be pumped or by stacking individual thin strips of the nanogel together with a proton source (e.g. electrodes) akin to the structure of biological muscles. Such a hierarchical structure could attain a fast, kHz response in a macroscopic gel. Our Letter reveals the dynamic interplay among conformation, solvent flow, and ionization in the swelling–collapse transition of nanogels. Our simulation setup enables the investigation of charge–structure–hydrodynamic coupling and can be utilized to investigate a wide range of solvated systems, including polyelectrolytes, proteins, and soft materials.

Acknowledgements.
This work was supported by the startup funds provided by the Whiting School of Engineering at JHU. The simulations were partially performed using the Advanced Research Computing at Hopkins (rockfish.jhu.edu), which is supported by the National Science Foundation (NSF) grant number OAC 1920103. We thank Erik Luijten, Hajime Tanaka, James D. Farrell, and Michael Falk for discussions and comments on the manuscript.

Appendix

Appendix A Effects of ion diffusion and protonation kinetics

Here we evaluate the effects of ion diffusion and protonation kinetics on the actuator performance. The timescale for ion diffusion across the nanogel studied is tdiff=L2/(2D)subscript𝑡diffsuperscript𝐿22𝐷t_{\textnormal{diff}}=L^{2}/(2D)italic_t start_POSTSUBSCRIPT diff end_POSTSUBSCRIPT = italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 2 italic_D ), which, for typical ion diffusivity D=nm2/ns𝐷superscriptnm2nsD=\textnormal{nm}^{2}/\textnormal{ns}italic_D = nm start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ns, is at least an order of magnitude faster than the fastest ram** rates considered in Figs. 2 and 3 (Lx/σ=NxNp=20subscript𝐿𝑥𝜎subscript𝑁𝑥subscript𝑁p20L_{x}/\sigma=N_{x}N_{\textnormal{p}}=20italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT / italic_σ = italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT p end_POSTSUBSCRIPT = 20). Since the timescales of collapse/expansion [Fig. 3(d)] and ion diffusion have the same scaling with the gel size (L2similar-toabsentsuperscript𝐿2\sim L^{2}∼ italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), we predict that ion diffusion across the gel is not expected to be a limiting process, even for bulk gels. The protonation dynamics is governed by typical weak acid dissociation rate kd3×106s1subscript𝑘d3superscript106superscripts1k_{\textnormal{d}}\approx 3\times 10^{6}~{}\textnormal{s}^{-1}italic_k start_POSTSUBSCRIPT d end_POSTSUBSCRIPT ≈ 3 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and association rate ka1011pHs1subscript𝑘asuperscript1011pHsuperscripts1k_{\textnormal{a}}\approx 10^{11-\textnormal{pH}}~{}\textnormal{s}^{-1}italic_k start_POSTSUBSCRIPT a end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT 11 - pH end_POSTSUPERSCRIPT s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT Kanzaki et al. (2014), which do not depend on the gel size. Since the response time due to solvent drag scales as L2superscript𝐿2L^{2}italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we conclude that the protonation dynamics can be a limiting factor only for nanoscale gels (L50nmless-than-or-similar-to𝐿50nmL\lesssim 50~{}\textnormal{nm}italic_L ≲ 50 nm) or extremely weak PEs where collapse occurs at high pH. For example, the actuation power in Figs. 2d and 3b are obtained at the high frequency limit allowable by protonation dynamics. The gel actuator is driven by protons and requires a proton source/sink, e.g., an electrode, and so ion/proton diffusion from the electrode to the gel could be another limiting factor. To achieve fast kHz or even MHz response time, the proton/source sink should be placed next to the gel, and it could itself be a soft electrode, e.g. a conjugated polyelectrolyte.

Appendix B Model limitations

Our model does not capture the orientational ordering of water near ions Van Der Vegt et al. (2016); Shi et al. (2023), which can enhance electrostatic coupling at small ion–ion distances. Moreover, ion and water diffusion are generally enhanced in polarizable compared to non-polarizable models Nguyen and Rick (2018). Thus, water ordering and polarizability are anticipated to enhance the transport and the response of charge to conformational changes, potentially improving the actuator’s performance.

References

  • Karg et al. (2019) Matthias Karg, Andrij Pich, Thomas Hellweg, Todd Hoare, L. Andrew Lyon, J. J. Crassous, Daisuke Suzuki, Rustam A. Gumerov, Stefanie Schneider, Igor. I. Potemkin,  and Walter Richtering, “Nanogels and microgels: From model colloids to applications, recent developments, and future trends,” Langmuir 35, 6231–6255 (2019).
  • van der Linden and Westerweel (2013) Heiko van der Linden and Jerry Westerweel, “Temperature-sensitive hydrogels,” in Encyclopedia of Microfluidics and Nanofluidics, edited by Dongqing Li (Springer US, Boston, MA, 2013) pp. 1–5.
  • Ge et al. (2021) Sijia Ge, Jiajia Li, Jian Geng, Shinian Liu, Hua Xu,  and Zhongze Gu, “Adjustable dual temperature-sensitive hydrogel based on a self-assembly cross-linking strategy with highly stretchable and healable properties,” Mater. Horiz. 8, 1189–1198 (2021).
  • Gupta et al. (2002) Piyush Gupta, Kavita Vermani,  and Sanjay Garg, “Hydrogels: From controlled release to pH-responsive drug delivery,” Drug Discov. Today 7, 569–579 (2002).
  • Glazer et al. (2012) P. J. Glazer, M. van Erp, A. Embrechts, S. G. Lemay,  and E. Mendes, “Role of pH gradients in the actuation of electro-responsive polyelectrolyte gels,” Soft Matt. 8, 4421–4426 (2012).
  • Han et al. (2020) Zilong Han, Peng Wang, Guoyong Mao, Tenghao Yin, Danming Zhong, Burebi Yiming, Xiaocheng Hu, Zheng Jia, Guodong Nian, Shaoxing Qu,  and Wei Yang, “Dual pH-responsive hydrogel actuator for lipophilic drug delivery,” ACS Appl. Mater. Interfaces 12, 12010–12017 (2020).
  • Shao et al. (2021) Zijian Shao, Shanshan Wu, Qian Zhang, Hui Xie, Tao Xiang,  and Shaobing Zhou, “Salt-responsive polyampholyte-based hydrogel actuators with gradient porous structures,” Polym. Chem. 12, 670–679 (2021).
  • Schreyer et al. (2000) H. Brett Schreyer, Nouvelle Gebhart, Kwang J. Kim,  and Mohsen Shahinpoor, “Electrical activation of artificial muscles containing polyacrylonitrile gel fibers,” Biomacromol. 1, 642–647 (2000).
  • Morales et al. (2014) Daniel Morales, Etienne Palleau, Michael D. Dickey,  and Orlin D. Velev, “Electro-actuated hydrogel walkers with dual responsive legs,” Soft Matt. 10, 1337–1348 (2014).
  • Li et al. (2020) Chuang Li, Garrett C. Lau, Hang Yuan, Aaveg Aggarwal, Victor Lopez Dominguez, Shuang** Liu, Hiroaki Sai, Liam C. Palmer, Nicholas A. Sather, Tyler J. Pearson, Danna E. Freedman, Pedram Khalili Amiri, Monica Olvera de la Cruz,  and Samuel I. Stupp, “Fast and programmable locomotion of hydrogel-metal hybrids under light and magnetic fields,” Sci. Robot. 5 (2020).
  • Li et al. (2022) Meng Li, Aniket Pal, Amirreza Aghakhani, Abdon Pena-Francesch,  and Metin Sitti, “Soft actuators for real-world applications,” Nature Reviews Materials 7, 235–249 (2022).
  • Quesada-Pérez et al. (2011) Manuel Quesada-Pérez, José Alberto Maroto-Centeno, Jacqueline Forcada,  and Roque Hidalgo-Alvarez, “Gel swelling theories: the classical formalism and recent approaches,” Soft Matt. 7, 10536–10547 (2011).
  • Dus̆ek and Patterson (1968) K. Dus̆ek and D. Patterson, “Transition in swollen polymer networks induced by intramolecular condensation,” J. Polym. Sci. Part A-2 6, 1209–1216 (1968).
  • Tanaka et al. (1980) Toyoichi Tanaka, David Fillmore, Shao-Tang Sun, Izumi Nishio, Gerald Swislow,  and Arati Shah, “Phase transitions in ionic gels,” Phys. Rev. Lett. 45, 1636–1639 (1980).
  • Yan and de Pablo (2003) Qiliang Yan and Juan J. de Pablo, “Monte Carlo simulation of a coarse-grained model of polyelectrolyte networks,” Phys. Rev. Lett. 91, 018301 (2003).
  • Curk and Luijten (2021) Tine Curk and Erik Luijten, “Charge regulation effects in nanoparticle self-assembly,” Phys. Rev. Lett. 126, 138003 (2021).
  • Curk et al. (2022) Tine Curk, Jiaxing Yuan,  and Erik Luijten, “Accelerated simulation method for charge regulation effects,” J. Chem. Phys. 156 (2022), 10.1063/5.0066432, 044122.
  • Polotsky et al. (2013) Alexey A. Polotsky, Felix A. Plamper,  and Oleg V. Borisov, “Collapse-to-swelling transitions in pH- and thermoresponsive microgels in aqueous dispersions: The thermodynamic theory,” Macromolecules 46, 8702–8709 (2013).
  • Zheng et al. (2023) Bin Zheng, Yael Avni, David Andelman,  and Rudolf Podgornik, “Charge regulation of polyelectrolyte gels: Swelling transition,” Macromolecules  (2023), 10.1021/acs.macromol.3c00609.
  • Keidel et al. (2018) Rico Keidel, Ali Ghavami, Dersy M Lugo, Gudrun Lotze, Otto Virtanen, Peter Beumers, Jan Skov Pedersen, Andre Bardow, Roland G Winkler,  and Walter Richtering, “Time-resolved structural evolution during the collapse of responsive hydrogels: The microgel-to-particle transition,” Sci. Adv. 4, eaao7086 (2018).
  • Dallari et al. (2024) Francesco Dallari, Irina Lokteva, Johannes Möller, Wojciech Roseker, Claudia Goy, Fabian Westermeier, Ulrike Boesenberg, Jörg Hallmann, Angel Rodriguez-Fernandez, Markus Scholz, et al., “Real-time swelling-collapse kinetics of nanogels driven by XFEL pulses,” Sci. Adv. 10, eadm7876 (2024).
  • Zhou et al. (2020) Wen Zhou, Guangzhao Yang, Xiaoyue Ni, Shanchao Diao, Chen Xie,  and Quli Fan, “Recent advances in crosslinked nanogel for multimodal imaging and cancer therapy,” Polymers 12, 1902 (2020).
  • Wang et al. (2021) Hao Wang, Lingfeng Gao, Taojian Fan, Chen Zhang, Bin Zhang, Omar A Al-Hartomy, Ahmed Al-Ghamdi, Swelm Wageh, Meng Qiu,  and Han Zhang, “Strategic design of intelligent-responsive nanogel carriers for cancer therapy,” ACS Appl. Mater. Interfaces 13, 54621–54647 (2021).
  • Asadian-Birjand et al. (2012) Mazdak Asadian-Birjand, Ana Sousa-Herves, Dirk Steinhilber, Julio César Cuggino,  and Marcelo Calderon, “Functional nanogels for biomedical applications,” Current medicinal chemistry 19, 5029–5043 (2012).
  • Podgornik (2018) Rudolf Podgornik, “General theory of charge regulation and surface differential capacitance,” J. Chem. Phys. 149, 104701 (2018).
  • Hockney and Eastwood (1988) R. W. Hockney and J. W. Eastwood, Computer Simulation Using Particles (IOP Publishing, Bristol, 1988).
  • Kikuchi et al. (2005) N. Kikuchi, J. F. Ryder, C. M. Pooley,  and J. M. Yeomans, “Kinetics of the polymer collapse transition: The role of hydrodynamics,” Phys. Rev. E 71, 061804 (2005).
  • Kamata et al. (2009) Kumiko Kamata, Takeaki Araki,  and Hajime Tanaka, “Hydrodynamic selection of the kinetic pathway of a polymer coil-globule transition,” Phys. Rev. Lett. 102, 108303 (2009).
  • Yuan et al. (2022) Jiaxing Yuan, Kyohei Takae,  and Hajime Tanaka, “Impact of inverse squeezing flow on the self-assembly of oppositely charged colloidal particles under electric field,” Phys. Rev. Lett. 129, 248001 (2022).
  • Yuan and Tanaka (2024) Jiaxing Yuan and Hajime Tanaka, “Hydrodynamic effects on the collapse kinetics of flexible polyelectrolytes,” Phys. Rev. Lett. 132, 038101 (2024).
  • Groot and Warren (1997) Robert D Groot and Patrick B Warren, “Dissipative particle dynamics: Bridging the gap between atomistic and mesoscopic simulation,” J. Chem. Phys. 107, 4423–4435 (1997).
  • Curk (2024) Tine Curk, “Dissipative particle dynamics for coarse-grained models,” The Journal of Chemical Physics 160, 174115 (2024).
  • (33) See Supplemental Material at http://link.aps.org/supplemental/xxx on the information about simulation details and additional characterization of pH-responsive nanogels.
  • Park and Kim (2020) Nuri Park and Jaeyun Kim, “Hydrogel-based artificial muscles: Overview and recent progress,” Adv. Intell. Syst. 2, 1900135 (2020).
  • Kang and Bae (2001) Seong Il Kang and You Han Bae, “pH-induced volume-phase transition of hydrogels containing sulfonamide side group by reversible crystal formation,” Macromolecules 34, 8173–8178 (2001).
  • Kanzaki et al. (2014) Yasushi Kanzaki, Koichi Tokuda,  and Stanley Bruckenstein, “Dissociation rates of weak acids using sinusoidal hydrodynamic modulated rotating disk electrode employing Koutecky-Levich equation,” J. Electrochem. Soc. 161, 770–779 (2014).
  • Ranatunga (1998) KW Ranatunga, “Temperature dependence of mechanical power output in mammalian (rat) skeletal muscle,” Exp. Physiol. 83, 371–376 (1998).
  • Mirfakhrai et al. (2007) Tissaphern Mirfakhrai, John D.W. Madden,  and Ray H. Baughman, “Polymer artificial muscles,” Materials Today 10, 30–38 (2007).
  • Kim et al. (2015) Youn Soo Kim, Mingjie Liu, Yasuhiro Ishida, Yasuo Ebina, Minoru Osada, Takayoshi Sasaki, Takaaki Hikima, Masaki Takata,  and Takuzo Aida, “Thermoresponsive actuation enabled by permittivity switching in an electrostatically anisotropic hydrogel,” Nature Mater. 14, 1002–1007 (2015).
  • Ma et al. (2020a) Yanfei Ma, Mutian Hua, Shuwang Wu, Yingjie Du, Xiaowei Pei, Xinyuan Zhu, Feng Zhou,  and Ximin He, “Bioinspired high-power-density strong contractile hydrogel by programmable elastic recoil,” Sci. Adv. 6 (2020a).
  • Tanaka and Fillmore (1979) Toyoichi Tanaka and David J Fillmore, “Kinetics of swelling of gels,” J. Chem. Phys. 70, 1214–1218 (1979).
  • Park et al. (2024) Chang Seo Park, Yong-Woo Kang, Hyeonuk Na,  and Jeong-Yun Sun, “Hydrogels for bioinspired soft robots,” Prog. Polym. Sci. , 101791 (2024).
  • Ma et al. (2020b) Yanfei Ma, Mutian Hua, Shuwang Wu, Yingjie Du, Xiaowei Pei, Xinyuan Zhu, Feng Zhou,  and Ximin He, “Bioinspired high-power-density strong contractile hydrogel by programmable elastic recoil,” Sci. Adv. 6, eabd2520 (2020b).
  • Van Der Vegt et al. (2016) Nico FA Van Der Vegt, Kristoffer Haldrup, Sylvie Roke, Junrong Zheng, Mikael Lund,  and Huib J Bakker, “Water-mediated ion pairing: Occurrence and relevance,” Chem. Rev. 116, 7626–7641 (2016).
  • Shi et al. (2023) Rui Shi, Anthony J Cooper,  and Hajime Tanaka, “Impact of hierarchical water dipole orderings on the dynamics of aqueous salt solutions,” Nature Comm. 14, 4616 (2023).
  • Nguyen and Rick (2018) Mary Nguyen and Steven W Rick, “The influence of polarizability and charge transfer on specific ion effects in the dynamics of aqueous salt solutions,” J. Chem. Phys. 148 (2018).