Introduction

Soft condensed matter (proteins, colloids or polymers) easily self-assembles into gels or amorphous soft solids with diverse structure and mechanics1,2,3,4,5. The nanoscale size of particles or aggregates makes these solids sensitive to thermal fluctuations, with spontaneous and thermally activated processes leading to rich microscopic dynamics. Besides affecting the time evolution, or aging, of the material properties at rest, such dynamics interplay with an imposed deformation and are hence crucial also for the mechanical response6,7,8,9,10. In the material at rest, thermal fluctuations can trigger ruptures of parts of the microscopic structure that are under tension or can promote coarsening and compaction of initially loosely packed domains, depending on the conditions under which the material initially solidified. The diffusion of particles or aggregates following such local reorganizations (often referred to as ‘micro-collapses’) is governed by a wide distribution of relaxation times, due to the disordered tortuous environment in which it takes place, that is, the microstructure of a soft solid, be it a thin fractal gel or a densely packed emulsion11,12. Therefore, localization, caging and larger-scale cooperative rearrangements in such complex environment will lead to slow, subdiffusive dynamics, similar to the microscopic dynamics close to a glass transition13,14.

Several quasi-elastic scattering experiments or numerical simulations that can access micro- and nanoscale rearrangements in soft matter have indeed confirmed that the dynamics in aging soft solids are slower than exponential. Measuring the time decay of the correlations of the density fluctuations or of the local displacements of particles or aggregates, the experiments find that time correlations follow a stretched exponential decay with β<1, akin to the slow cooperative dynamics of supercooled liquids and glasses, in a wide range of gels and other soft solids15,16,17,18,19. Nevertheless, in the past few years there has been emerging evidence, through a growing body of similar experiments, that microscopic dynamics in the aging of soft materials can be instead faster than exponential. In a wide range of soft amorphous solids including colloidal gels, biopolymer networks, foams and densely packed microgels, time correlations measured via quasi-elastic scattering or other time-resolved spectroscopy techniques appear to decay as with β>1 (and in most cases 1.3≤β≤1.5)20,21,22,23,24,25,26,27,28,29,33. Therefore, we expect our general picture to have a much wider relevance. Second, the physical mechanisms we propose help rationalize several experimental observations in very different materials, ranging from biologically relevant soft solids to metallic glasses31,45,46,50. In a quasi-equilibrium scenario, enthalpic and thermal degrees of freedom may still couple and stress correlations decay relatively fast. When the material is deeply quenched and jammed, instead, recovering the coupling between the distinct degrees of freedom and restoring equilibrium will require timescales well beyond the ones accessible in typical experiments or simulations. The result will be intermittent dynamics and compressed exponential relaxations. The competition between Brownian motion and elastic effects through the relaxation of internal stresses illustrated in this work suggests different scenarios for the energy landscape underlying the aging of soft jammed materials. When thermal fluctuations screen the long-range elastic strain transmission, microscopic rearrangements may open paths to deeper and deeper local minima in a rugged energy landscape. Compressed exponential dynamics, instead, evoke the presence of flat regions and huge barriers, with the possibility of intermittent dynamics, abrupt rearrangements and avalanches34,35. Investigating how such different dynamical processes couple with imposed deformations will provide a new rationale, and have important implications, for designing mechanics, rheology and material performances.

Methods

Numerical model and viscoelastic parameters

The particles in the model gel interact through a potential composed of two terms. The first force contribution derives from a Lennard–Jones-like potential of the form:

The second contribution confers an angular rigidity to the inter particles bonds and takes the form:

The strength of the interaction is controlled by Λ(r) and vanishes over two particles diameters:

where Θ denotes the Heaviside function. The evolution of the gel over time is obtained by solving the following Langevin equation for each particle:

where σ is the particle diameter, ξ(t) is a random white noise that models the thermal fluctuations and is related to the friction coefficient ηf by means of its variance . To be in the overdamped limit of the dynamics ηf is set to 10, and the timestep δt used for the numerical integration is δt=0.005. The parameters of the potential are chosen such that the disordered thin percolating network starts to self assemble at kBT/ = 0.05. One convenient choice to achieve this configuration is given by this set of parameters: A=6.27, a=0.85, B=67.27, θ=65° and w=0.3. The system is composed of N=62,500 particles in a cubic simulation box of a size L=76.43σ with periodic boundary conditions, the number density N/L3 is fixed at 0.14, which corresponds approximatively to a volume fraction of 7%. All initial gel configurations are the same, prepared with the protocol described in ref. 40, which consists in starting from a gas configuration (kBT/ = 0.5) and letting the gel self-assemble upon slow cooling down to kBT/ = 0.05. We then quench this gel configuration by running a simulation with the dissipative dynamics until the kinetic energy drops to zero(10−24). All simulations have been performed using a version of LAMMPS suitably modified by us52.

Stress calculation and cutting strategy

We let the initial gel configuration evolve with equation (4) for each of the different values of kBT/ considered here, while using the following procedure to cut network connections. At each timestep, we characterize the state of stress of a gel configuration by computing the virial stresses as , where the Greek subscripts stand for the Cartesian components x, y, z and represents the contribution to the stress tensor of all the interactions involving the particle i and V is the total volume of the simulation box52. contains, for each particle, the contributions of the two-body and the three-body forces evenly distributed among the particles that participate in them:

The first term on the RHS denotes the contribution of the two-body interaction, where the sum runs over all the N2 pair of interactions that involve the particle i. (ri, Fi) and (r′, F′) denote, respectively, the position and the forces on the two interacting particles. The second term indicates the three-body interactions involving the particle i. We consider a coarse-graining volume Ωcg centered around the point of interest r and containing around 9–10 particles on average, and define the local coarse-grained stress based on the per-particle virial contribution as . For a typical starting configuration of the gel, the local normal stress reflect the heterogeneity of the structure and tend to be higher around the nodes, due to the topological frustration of the network. We consider that breaking of network connections underlying the aging of the gel is more prone to happen in the regions where local stresses tend to be higher, as found also in refs 40, 41. Hence, to mimic the aging in the molecular dynamics simulations we scan the whole structure of the gel and remove one of the bonds (by turning off the well in ) whose contribution to the local normal stress is the largest (prevalently bonds between particles belonging to the network nodes). As the simulation proceeds, local internal stresses redistribute in the aging structure of the gel and the locations of more probable connection rupture (as well as their number) change over time. All simulations discussed here have been performed with a rate , corresponding to removing only ∼5% of the total network connections over the whole simulation time window. We have run several tests varying the removal rate in order to verify that changing the rate (having kept all other parameters constant) does not modify our outcomes and the emerging physical picture. Overall, varying Γ over nearly two orders of magnitudes, we recover the same results, as long as the τr=1/Γ between two rupture events allows for at least partial stress relaxation (see Fig. 1).

Stress autocorrelation function

The autocorrelation function of the stress fluctuations is computed as follows: , where fδσ is the auto-covariance and takes the form and . We have computed the autocorrelation function of the stress fluctuations over partial time series and over the whole duration of the simulations (see Fig. 5 and Supplementary Fig. 2).

Fitting procedure for the intermediate scattering functions

To extract the β exponent and the structural relaxation time τq, we fit the last decay of the coherent scattering function F(q,t)tw=0 simultaneously with the incoherent scattering function , defined as . This observable is less noisy and has the same exponent as the coherent scattering function shown in Fig. 3 (see also Supplementary Fig. 3). Having fitted Fs(q,t) with a stretched exponential type of the form and having extracted the exponent β, such exponent is used to initiate the fit of F(q,t) by locking this parameter and letting all the others free. This procedure helps us to improve the quality of the fit used to extract the relaxation time τq from the coherent scattering.

Data availability

The authors declare that all data supporting the findings of this study are available within the article and its Supplementary Information Files. All raw data can be accessed on request to the authors.

Additional information

How to cite this article: Bouzid, M. et al. Elastically driven intermittent microscopic dynamics in soft solids. Nat. Commun. 8, 15846 doi: 10.1038/ncomms15846 (2017).

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.