Density and geometry of excitations in supercooled liquids up to the activation energy

Wencheng Ji Department of Physics of Complex Systems, Weizmann Institute of Science, Rehovot, 234 Hertzl St., Israel    Massimo Pica Ciamarra [email protected] Division of Physics and Applied Physics, School of Physical and Mathematical Sciences, Nanyang Technological University, 21 Nanyang Link, 637371, Singapore Consiglio Nazionale delle Ricerce, CNR-SPIN, Napoli, I-80126, Italy    Matthieu Wyart [email protected] Institute of Physics, École Polytechnique Fédérale de Lausanne, Lausanne, CH-1015, Switzerland
Abstract

We introduce an algorithm to uncover the activated particle rearrangements, or excitations, regulating structural relaxation in glasses at much higher energies than previously achieved. We use it to investigate the density and geometric properties of excitations in a model system. We find that the density of excitations behaves as a shifted power-law, and confirm that this shift accounts for the increase in the activation energy controlling the relaxation dynamics. Remarkably, we find that excitations comprise a core whose properties, including the displacement of the particle moving the most, scale as a power-law of their activation energy and do not depend on temperature. Excitations also present an outer deformation field that depends on the material stability and, hence, on temperature. Our analysis suggests that while excitations suppress the transition of dynamical arrest predicted by mean-field theories, they are strongly influenced by it.

In low-temperature glasses, elementary excitations are two-level systems, groups of particles that can tunnel between two states [1, 2, 3, 4, 5, 6]. At much higher temperatures near the glass transition Tgsubscript𝑇𝑔T_{g}italic_T start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, structural relaxation occurs on a time scale τ=τ0exp(Ea/T)𝜏subscript𝜏0subscript𝐸𝑎𝑇\tau=\tau_{0}\exp(E_{a}/T)italic_τ = italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_exp ( italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT / italic_T ), where τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is microscopic time and Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is an activation energy that grows on cooling in fragile liquids [7]. Thermally activated local rearrangements of a few particles are important in this temperature range as well, as they contribute to structural relaxation [8, 9, 10, 11]. Such rearrangements thus play a vital role in theories of the glass transition  [12, 13, 14, 15]. In Kinetically Constrained Models, local rearrangements are defects that can diffuse and interact to relax the system [16, 17]. In mean field theories, local rearrangements correspond to the string-like [18, 19, 20, 21] ‘hop** processes’ through which finite-dimensional systems relax below the dynamical or mode coupling temperature Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, where the dynamics would halt in infinite dimensions where these processes are absent [22, 23]. Finally, in elastic models of the glass transition [24, 25, 26, 27, 28], the energy of local excitations directly determines the activation energy Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT. Theoretically, distinct ideas have been proposed to understand the geometry of local rearrangements, including entropic considerations [29] or the existence of defects around a hexatic phase in two dimensions [17]. Alternatively, building on the notion of a length scale diverging [30, 31, 32] at the dynamical transition Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, Ref [33] derived relationships between geometric and energetic properties of the excitations with the minimal energy.

Differentiating between different scenarios for the geometry of these excitations requires measuring their density of states N(E)𝑁𝐸N(E)italic_N ( italic_E ) and geometry across a broad energy spectrum, a challenge that remains unresolved. Indeed, potential energy landscape studies [34, 35] access the consecutive excitations activated in a liquid during its relations but cannot provide the density of states of excitations. Studies conducted on the few lowest-energy excitations [36, 37, 38, 6, 39], are pertinent to the plastic and quantum properties of glasses and not directly relevant to the glass transition, which typically involves much higher energy rearrangements. We recently developed SEER [40], an algorithm based on thermal exploration that allowed us to measure N(E)𝑁𝐸N(E)italic_N ( italic_E ), albeit only for energy notably smaller than the activation energy Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT. However, this study did not consider the geometry of excitations. Moreover, excitations with activation energy up to Tlog(τ/τ0)𝑇𝜏subscript𝜏0T\log(\tau/\tau_{0})italic_T roman_log ( italic_τ / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) have a significant probability to be triggered dynamically, and can then act as transition states enabling relaxation [34]. The excitations’ density of state and geometric properties in such a broad energy range have yet to be studied.

In this Letter, we perform such a study by introducing an algorithm to identify excitations via mechanical perturbations. This algorithm offers fast computation and enables access to a wide range of energies, surpassing the activation energy of the model under consideration. Our findings reveal two key insights: (i) We observe that the density of states N(E)𝑁𝐸N(E)italic_N ( italic_E ) approximately follows N(E)(EEg(T))aproportional-to𝑁𝐸superscript𝐸subscript𝐸𝑔𝑇𝑎N(E)\propto(E-E_{g}(T))^{a}italic_N ( italic_E ) ∝ ( italic_E - italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_T ) ) start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT up to the activation energy. Additionally, we confirm that the variation in Eg(T)subscript𝐸𝑔𝑇E_{g}(T)italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_T ) predicts the change in activation energy [40]. (ii) Remarkably, certain geometric properties, such as the displacement δ𝛿\deltaitalic_δ of the most-moving particle or the probability of displaying string-like motion, only depend on their energy E𝐸Eitalic_E. However, other properties, like the volume of an excitation, depend on both energy and temperature. We reconcile these observations by considering that excitations possess a core solely governed by their energy, which then influences the surrounding medium on a scale determined by material stability and, consequently, temperature. Overall, our results suggest that the hop** processes that suppress the mean-field dynamical transition are very much affected by it.

Refer to caption
Figure 1: Illustration of the ASEER algorithm. We detect excitations by enforcing a dipolar force through a spring connecting adjacent particles of energy-minimized configurations. We gradually increase the spring rest length until an instability occurs. We then remove the spring and minimize the energy, bringing the system to a novel energy minimum. The minimum energy pathway going from one state to the other is analyzed in the absence of the spring.

ASEER – We propose an athermal algorithm, the Athermal Systematic Excitation ExtRaction or ‘ASEER’, to uncover the excitations associated with a reference inherent structure (IS0). This protocol builds on the idea that changing the topology of the Voronoi neighbors induces an excitation, in analogy with the triggering of T1 transition in two spatial dimensions, see e.g. Ref. [27, 28, 41]. To uncover an excitation, we increase the separation between two adjacent (à la Voronoi) particles i𝑖iitalic_i and j𝑗jitalic_j by modifying the energy functional via the addition of an elastic spring, Espring(Δr)=k[(rij0+Δr)rij]2subscript𝐸springΔ𝑟𝑘superscriptdelimited-[]superscriptsubscript𝑟𝑖𝑗0Δ𝑟subscript𝑟𝑖𝑗2E_{\rm spring}(\Delta r)=k[(r_{ij}^{0}+\Delta r)-r_{ij}]^{2}italic_E start_POSTSUBSCRIPT roman_spring end_POSTSUBSCRIPT ( roman_Δ italic_r ) = italic_k [ ( italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + roman_Δ italic_r ) - italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, rij0superscriptsubscript𝑟𝑖𝑗0r_{ij}^{0}italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT being the distance between the particles in IS0. We have checked that the value of k𝑘kitalic_k is not critical and empirically fix it so that the spring is never compressed by more than 5%percent55\%5 %. We slowly increase ΔrΔ𝑟\Delta rroman_Δ italic_r while continuously minimizing the energy to keep the system in a minimum of the expanded energy functional U({𝐫𝐢})+Espring(Δr)𝑈subscript𝐫𝐢subscript𝐸springΔ𝑟U(\{{\bf r_{i}}\})+E_{\rm spring}(\Delta r)italic_U ( { bold_r start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT } ) + italic_E start_POSTSUBSCRIPT roman_spring end_POSTSUBSCRIPT ( roman_Δ italic_r ). The ΔrΔ𝑟\Delta rroman_Δ italic_r dependence of the spring energy (or of the total one) comprises smooth elastic branches punctuated by sudden drops corresponding to irreversible rearrangements. We focus on the first plastic event identified via a standard thresholding approach. When this event occurs, we remove the spring and minimize the energy again, potentially leading the system to a new IS. This approach is illustrated in Fig.1 and involves constraining one degree of freedom. We investigate the minimum energy path connecting IS0, and each uncovered IS via the nudge-elastic-band (NEB) method [42] to estimate the energy ESaddlesubscript𝐸SaddleE_{\rm Saddle}italic_E start_POSTSUBSCRIPT roman_Saddle end_POSTSUBSCRIPT of the saddle point separating the considered ISs. If not stated otherwise, when the minimum energy path traverses additional ISs (20%similar-to-or-equalsabsentpercent20\simeq 20\%≃ 20 % of cases in the considered system), we redefine IS as the first encountered IS and repeat the NEB analysis. This results in a catalog of unique (we ensure each IS appears once) excitations, each characterized by its energy barrier E=ESaddleEIS0𝐸subscript𝐸Saddlesubscript𝐸subscriptIS0E=E_{\rm Saddle}-E_{\rm IS_{0}}italic_E = italic_E start_POSTSUBSCRIPT roman_Saddle end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT roman_IS start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT and displacement field 𝐝𝐫𝐝𝐫\bf{dr}bold_dr. This algorithm leads to large catalogs of excitations, as it uncovers from 1.21.21.21.2 to 2222 excitations per particle on cooling. As demonstrated below, at low energies, ASEER matches previous algorithms systematically searching excitations via thermal fluctuations, while it performs much better at energies of the order of the activation energy Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT.

Refer to caption
Figure 2: (a) Density of excitations N(E)𝑁𝐸N(E)italic_N ( italic_E ), normalized by the system size 𝒩𝒩\mathcal{N}caligraphic_N, and its cumulative distribution F(E)𝐹𝐸F(E)italic_F ( italic_E ) (inset). The solid dots mark the values of the activation energy [40], Ea(T)=Tlog(τ/τ0)subscript𝐸𝑎𝑇𝑇𝜏subscript𝜏0E_{a}(T)=T\log\left(\tau/\tau_{0}\right)italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_T ) = italic_T roman_log ( italic_τ / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), and the cross-shaped line has been obtained with the SEER algorithm at T=0.5𝑇0.5T=0.5italic_T = 0.5. (b) The N(E)𝑁𝐸N(E)italic_N ( italic_E ) curves collapse when the energy is shifted by Eg(T)subscript𝐸𝑔𝑇E_{g}(T)italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_T ). The thick green curves is N(E)g1×(EEg)2.7±0.1𝑁𝐸subscript𝑔1superscript𝐸subscript𝐸𝑔plus-or-minus2.70.1N(E)\!\approx g_{1}\!\times\!(E\!-\!E_{g})^{2.7\pm 0.1}italic_N ( italic_E ) ≈ italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT × ( italic_E - italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2.7 ± 0.1 end_POSTSUPERSCRIPT, where g1=(4.5±0.5)×104subscript𝑔1plus-or-minus4.50.5superscript104g_{1}\!=\!\!(4.5\pm 0.5)\!\times\!10^{-4}italic_g start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = ( 4.5 ± 0.5 ) × 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT. (c) The shift of Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT on cooling matches the shift in activation energy evaluated via SEER [40]. (d) The increase in Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT matches the increase in activation energy measured from the relaxation dynamics, with T=0.5𝑇0.5T=0.5italic_T = 0.5 an arbitrary reference temperature.

Density of states – We applied ASEER to a polydisperse three-dimensional system of N=2000𝑁2000N=2000italic_N = 2000 soft repulsive particles [43] that can be equilibrated up to experimentally comparable temperatures through the ‘swap’ algorithm [44, 45, 46, 47]. In recent work [40], we determined this model’s relaxation time τ(T)𝜏𝑇\tau(T)italic_τ ( italic_T ) and microscopic timescale τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and hence the temperature dependence of its activation energy, Ea(T)=Tlog(τ/τ0)subscript𝐸𝑎𝑇𝑇𝜏subscript𝜏0E_{a}(T)=T\log(\tau/\tau_{0})italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_T ) = italic_T roman_log ( italic_τ / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). In the Supplemental Material (SM) [48], we provide numerical details on the numerical model and the measure of the relaxation time.

Fig. 2(a) illustrates the energy E𝐸Eitalic_E dependence of the density of excitations N(E)𝑁𝐸N(E)italic_N ( italic_E ), and its cumulative F(E)𝐹𝐸F(E)italic_F ( italic_E )(inset). At each temperature T𝑇Titalic_T, we average over ten samples. Notably, N(E)𝑁𝐸N(E)italic_N ( italic_E ) and F(E)𝐹𝐸F(E)italic_F ( italic_E ) approximately shifts towards higher energy values as T𝑇Titalic_T decreases: the energy of local barriers grows under cooling. Indeed, before its maximum N(E)𝑁𝐸N(E)italic_N ( italic_E ) is well-described by:

N(E)(EEg(T))a.proportional-to𝑁𝐸superscript𝐸subscript𝐸𝑔𝑇𝑎N(E)\propto(E\!-\!E_{g}(T))^{a}.italic_N ( italic_E ) ∝ ( italic_E - italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_T ) ) start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT . (1)

with a2.7𝑎2.7a\approx 2.7italic_a ≈ 2.7, as demonstrated by the data collapse in Fig. 2(b). The characteristic energy gap Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT characterizing the system’s stability increases on cooling, as shown in Fig. 2(c). Fig. 2(d) shows that the variation of Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT on cooling is consistent with that of the activation energy, indicating that local barriers control the dynamics in this liquid. For the two lowest-temperatures (open circle), Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is extracted from values of τ𝜏\tauitalic_τ estimated through the time-temperature superposition principle [49, 48, 40]. The increase of a gap Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT is reminiscent of the mean-field prediction according to which, for T<Tc𝑇subscript𝑇𝑐T<T_{c}italic_T < italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the density of vibrational modes is gapped up to a frequency ωminsubscript𝜔\omega_{\min}italic_ω start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT indeed increasing on cooling [50].

In Fig. 2(a) and (c), we also present data obtained previously using the SEER algorithm [40]. While both algorithms yield consistent results for the variation of Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (and of other quantities in Fig. S1 of [48]), ASEER notably detects excitation on the scale Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, which can be activated on the relaxation time scale. Each algorithm possesses distinct advantages: SEER, reliant on thermal cycles, uncovers excitations with the smallest E𝐸Eitalic_E, making it ideal for obtaining accurate statistics at lower E𝐸Eitalic_E values where N(E)𝑁𝐸N(E)italic_N ( italic_E ) is limited. Conversely, the mechanical ASEER algorithm accesses a wide spectrum of activation energies. Noticing that the algorithms give consistent results in the energy range accessed by both (Fig. S1), we integrate them in the subsequent analysis to acquire high-quality statistics across a broad energy range.

Architecture of lowest-energy excitations–

Refer to caption
Figure 3: Dependence of (a) volume, (b) largest particle displacement, and (c) squared norm of the displacement field on the excitation’s activation energy, for various temperatures. The green stars refer to the excitations with the smallest activation energy. In all panels, we plot median values (see Fig. S2 in [48] for mean values). (d) The data in (c) are collapsed by a model postulating that the excitation core induces a T𝑇Titalic_T-dependent far-field displacement.

We analyze the geometrical properties of each excitation by its displacement field, 𝐝𝐫𝐝𝐫{\bf dr}bold_dr, focusing on: (i) the characteristic number of involved particles, estimated by the participation ratio VNPr=(𝐝𝐫i2)2/(𝐝𝐫i4)𝑉𝑁subscript𝑃𝑟superscriptsuperscriptsubscript𝐝𝐫𝑖22superscriptsubscript𝐝𝐫𝑖4V\equiv NP_{r}=(\sum{\bf dr}_{i}^{2})^{2}/(\sum{\bf dr}_{i}^{4})italic_V ≡ italic_N italic_P start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = ( ∑ bold_dr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( ∑ bold_dr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ), where the sums run over all particles i𝑖iitalic_i; (ii) the squared norm of the displacement field, |dr|2=i𝐝𝐫i2superscript𝑑𝑟2subscript𝑖superscriptsubscript𝐝𝐫𝑖2|dr|^{2}\!=\!\sum_{i}{\bf dr}_{i}^{2}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bold_dr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT; (iii) the maximum particle displacement, δ=maxi𝐝𝐫i𝛿subscript𝑖normsubscript𝐝𝐫𝑖\delta=\max_{i}||{\bf dr}_{i}||italic_δ = roman_max start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | | bold_dr start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | |.

By elaborating on the mean-field scenario [30, 31, 32], Ref. [33] predicted that for the excitations with minimal energy, associated with modes near the gap:

Vmin(T)1δmin(T)1|drmin(T)|21Emin(T)13.similar-tosubscript𝑉𝑇1subscript𝛿𝑇similar-to1superscript𝑑subscript𝑟𝑇2similar-to1subscript𝐸superscript𝑇13V_{\min}(T)\sim\frac{1}{\delta_{\min}(T)}\sim\frac{1}{|dr_{\min}(T)|^{2}}\sim% \frac{1}{{E_{\min}(T)}^{\frac{1}{3}}}.italic_V start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) ∼ divide start_ARG 1 end_ARG start_ARG italic_δ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) end_ARG ∼ divide start_ARG 1 end_ARG start_ARG | italic_d italic_r start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∼ divide start_ARG 1 end_ARG start_ARG italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_POSTSUPERSCRIPT end_ARG . (2)

These excitations are also associated with a length scale min(T)Vmin(T)similar-tosubscript𝑇subscript𝑉𝑇\ell_{\min}(T)\sim\sqrt{V_{\min}(T)}roman_ℓ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) ∼ square-root start_ARG italic_V start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) end_ARG, which is empirically known to characterize the linear response to an imposed dipole  [51, 52]. Ref. [33] verified the relationships between Vminsubscript𝑉V_{\min}italic_V start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, δmin(T)subscript𝛿𝑇\delta_{\min}(T)italic_δ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) and |drmin(T)|2superscript𝑑subscript𝑟𝑇2|dr_{\min}(T)|^{2}| italic_d italic_r start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, but the activation energy Eminsubscript𝐸E_{\min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT was not measured; instead a proxy corresponding to the energy difference between the two IS was used. We validate the scaling of the excitation architecture with Eminsubscript𝐸E_{\min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT in Fig. 3(a), (b) and (c) (green stars) by investigating, at each temperature, the features of the lowest-energy excitation of 1000 systems.

Note that we exclude string-like excitations from the geometrical analysis (they are defined as excitations involving the exchange of particles; see below). In general, including them only affects results for the largest energies considered, as shown in Fig. S4 [48].

Architecture of high-energy excitations – Figure 3 illustrates our main results, how the excitations’ geometric properties depend on E𝐸Eitalic_E and T𝑇Titalic_T, for E>Emin𝐸subscript𝐸E>E_{\min}italic_E > italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT. Panel (b) shows a remarkable result: the largest displacement in an excitation depends on the excitation energy but not on its temperature so that δE1/3proportional-to𝛿superscript𝐸13\delta\propto E^{1/3}italic_δ ∝ italic_E start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT. By contrast, panels (a) and (c) demonstrate that, at fixed energy, the volume of an excitation and its squared norm increase with temperature. These facts can be visualized by considering the two-dimensional projection of the displacement field of excitations of similar energy E10𝐸10E\approx 10italic_E ≈ 10 in systems having different temperatures, T=0.3𝑇0.3T=0.3italic_T = 0.3 and T=0.5𝑇0.5T=0.5italic_T = 0.5. Fig. 4 (a) and (b) show that these excitations have a comparable maximal magnitude at their center, but differ in the far field.

Refer to caption
Figure 4: Representative 2D-projections of the displacement fields of excitations E10𝐸10E\approx 10italic_E ≈ 10 and δ1𝛿1\delta\approx 1italic_δ ≈ 1 at temperatures T=0.3𝑇0.3T=0.3italic_T = 0.3 (a) and T=0.5𝑇0.5T=0.5italic_T = 0.5 (b). The displacement fields are scaled by a factor of three for visualization purposes. Response fields induced by a dipole force acting on neighboring particles (brown dots) at T=0.3𝑇0.3T=0.3italic_T = 0.3 (c) and T=0.5𝑇0.5T=0.5italic_T = 0.5 (d).

We rationalize these observations by assuming that excitations consist of (i) a core that depends only on energy E𝐸Eitalic_E and (ii) a far-field elastic response triggered by the core. In this view, the norm square of an excitation consists of two parts |dr|2(T,E)=|dr|core2+|dr|f.f.2superscript𝑑𝑟2𝑇𝐸subscriptsuperscript𝑑𝑟2coresubscriptsuperscript𝑑𝑟2formulae-sequenceff|dr|^{2}(T,E)=|dr|^{2}_{\rm core}+|dr|^{2}_{\rm{f.f.}}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_T , italic_E ) = | italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_core end_POSTSUBSCRIPT + | italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_f . roman_f . end_POSTSUBSCRIPT, which we now estimate.

Assumption (i) implies that the core is the same if E=Emin𝐸subscript𝐸E=E_{\min}italic_E = italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT or E>Emin𝐸subscript𝐸E>E_{\min}italic_E > italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, as long as the energy is the same. Together with Eq. 2, it thus leads to a maximal displacement of the core, and therefore of the whole excitation, behaving as δE1/3similar-to𝛿superscript𝐸13\delta\sim E^{1/3}italic_δ ∼ italic_E start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT and a square norm of the core |dr|core2(E)δ2(E)V(E)=c0E1/3proportional-tosubscriptsuperscript𝑑𝑟2core𝐸superscript𝛿2𝐸𝑉𝐸subscript𝑐0superscript𝐸13|dr|^{2}_{\rm core}(E)\propto\delta^{2}(E)V(E)=c_{0}E^{1/3}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_core end_POSTSUBSCRIPT ( italic_E ) ∝ italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_E ) italic_V ( italic_E ) = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT.

If EEminmuch-greater-than𝐸subscript𝐸E\gg E_{\min}italic_E ≫ italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, we expect this core to act as a local strain, which generically triggers a dipolar linear elastic response in the far field [53]. It is known that the response to a unit dipole is more extended at large rather than small temperatures [51, 52], as we explicitly show in Fig. 4 (c) and (d). Quantitatively, following [51, 33] we expect the volume of a unit dipole response to vary with temperature as Vmin(T)subscript𝑉𝑇V_{\min}(T)italic_V start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ), of order Emin(T)1/3similar-toabsentsubscript𝐸superscript𝑇13\sim E_{\min}(T)^{-1/3}∼ italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT according to Eq. 2. Moreover, we expect the norm of the dipole response to be proportional to the norm of the core itself. Putting these two effects together, we obtain |dr|f.f.2f(E/Emin)|dr|core2(E)Vmin(T)f(E/Emin)(E/Emin)1/3proportional-tosubscriptsuperscript𝑑𝑟2formulae-sequenceff𝑓𝐸subscript𝐸subscriptsuperscript𝑑𝑟2core𝐸subscript𝑉𝑇proportional-to𝑓𝐸subscript𝐸superscript𝐸subscript𝐸13|dr|^{2}_{\rm f.f.}\propto f(E/E_{\min})|dr|^{2}_{\rm core}(E)V_{\min}(T)% \propto f(E/E_{\min})(E/E_{\min})^{1/3}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_f . roman_f . end_POSTSUBSCRIPT ∝ italic_f ( italic_E / italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) | italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_core end_POSTSUBSCRIPT ( italic_E ) italic_V start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) ∝ italic_f ( italic_E / italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) ( italic_E / italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT. Here the function f(x)𝑓𝑥f(x)italic_f ( italic_x ), with f(1)=0𝑓10f(1)=0italic_f ( 1 ) = 0 and limxf(x)=C>0subscript𝑥𝑓𝑥𝐶0\lim_{x\rightarrow\infty}f(x)=C>0roman_lim start_POSTSUBSCRIPT italic_x → ∞ end_POSTSUBSCRIPT italic_f ( italic_x ) = italic_C > 0, accounts for the fact that only high-energy excitations (EEminmuch-greater-than𝐸subscript𝐸E\gg E_{\min}italic_E ≫ italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT) produce a far-field response that differs from the core. Overall, we find:

|dr|2(T,E)=c0E1/3+f(EEmin)(EEmin)1/3.superscript𝑑𝑟2𝑇𝐸subscript𝑐0superscript𝐸13𝑓𝐸subscript𝐸superscript𝐸subscript𝐸13\displaystyle|dr|^{2}(T,E)=c_{0}E^{1/3}\!+\!f\!\left(\!\frac{E}{E_{\min}}\!% \right)\left(\frac{E}{E_{\min}}\right)^{1/3}.| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_T , italic_E ) = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_E start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT + italic_f ( divide start_ARG italic_E end_ARG start_ARG italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_E end_ARG start_ARG italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT . (3)

Fig. 3(d) validates this theoretical prediction for a natural choice of interpolating function f(x)(x1)/(x+1)proportional-to𝑓𝑥𝑥1𝑥1f(x)\propto(x-1)/(x+1)italic_f ( italic_x ) ∝ ( italic_x - 1 ) / ( italic_x + 1 ). In Fig. S3 of SM, we show that Eq. 3 also holds if Eminsubscript𝐸E_{\min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT is replaced by E5subscript𝐸5E_{5}italic_E start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT, E10subscript𝐸10E_{10}italic_E start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT (respectively the median of the five or ten lowest energy excitation) or Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT.

String-like excitations (SLEs) – Numerical simulations [54, 55, 39] and recent experiments [56] show string-like particle motion in low-temperature structural relaxation. String-like rearrangements involve one or more particles swap** positions in the excitation core, with displacements comparable to inter-particle separation [33]. We thus expect that (i) High-energy excitations are more likely string-like due to their larger displacements. (ii) The probability of a string-like excitation is T𝑇Titalic_T-independent as the core properties are T𝑇Titalic_T-independent.

We identify SLEs by assuming that particle i𝑖iitalic_i ends up in the position originally occupied by particle ji𝑗𝑖j\neq iitalic_j ≠ italic_i if |𝐫j𝐫i0|<Δsubscriptsuperscript𝐫𝑗superscriptsubscript𝐫𝑖0Δ|{\bf{r}}^{*}_{j}\!-\!{\bf{r}}_{i}^{0}|<\Delta| bold_r start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT | < roman_Δ, 𝐫i0superscriptsubscript𝐫𝑖0{\bf{r}}_{i}^{0}bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and 𝐫jsubscriptsuperscript𝐫𝑗{\bf{r}}^{*}_{j}bold_r start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT being the positions in the initial and final configurations. The choice of ΔΔ\Deltaroman_Δ is not critical as long as its value corresponds to a small fraction of the inter-particle distance. Here, we fix Δ=0.1Δ0.1\Delta=0.1roman_Δ = 0.1. The number of particles involved in a string defines a string length np=ij,jiθ(0.1|𝐫j𝐫i0|)subscript𝑛𝑝subscript𝑖subscript𝑗𝑗𝑖𝜃0.1subscriptsuperscript𝐫𝑗superscriptsubscript𝐫𝑖0n_{p}=\sum_{i}\sum_{j,j\neq i}\theta\left(0.1-|{\bf{r}}^{*}_{j}\!-\!{\bf{r}}_{% i}^{0}|\right)italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j , italic_j ≠ italic_i end_POSTSUBSCRIPT italic_θ ( 0.1 - | bold_r start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT | ). An excitation is string-like provided np1subscript𝑛𝑝1n_{p}\geq 1italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≥ 1. An example string-like excitation is in Animation 1 [48]. Fig. (5)(a) shows that an excitation is a string with a T𝑇Titalic_T-independent probability Fs=θ(np1)subscript𝐹𝑠delimited-⟨⟩𝜃subscript𝑛𝑝1F_{s}=\langle\theta(n_{p}-1)\rangleitalic_F start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ⟨ italic_θ ( italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - 1 ) ⟩ that increases with E𝐸Eitalic_E, consistently with our expectations. The string length similarly increases with E𝐸Eitalic_E and is approximately T𝑇Titalic_T-independent, as we illustrate in Fig. (5)(b).

We delve deeper into the dependence of an excitation’s spatial and energetic properties on the string length. Figure (5)(c) demonstrates how the energy change |EfEi|subscript𝐸𝑓subscript𝐸𝑖|E_{f}-E_{i}|| italic_E start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | between the initial and final inherent structures varies with E𝐸Eitalic_E, considering strings of different lengths at T=0.4𝑇0.4T=0.4italic_T = 0.4. Notably, only a few excitations with np=0,1subscript𝑛𝑝01n_{p}=0,1italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 0 , 1 exhibit |EfEi|0similar-to-or-equalssubscript𝐸𝑓subscript𝐸𝑖0|E_{f}-E_{i}|\simeq 0| italic_E start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ≃ 0, while this trend becomes more prevalent for larger npsubscript𝑛𝑝n_{p}italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, particularly for np=2subscript𝑛𝑝2n_{p}=2italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 2. This phenomenon arises because many excitations with greater npsubscript𝑛𝑝n_{p}italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT correspond to closed strings, as depicted in Fig. (5)(d), and consequently have minimal impact on the system’s structure and energy. By promoting particle motion without inducing substantial structural relaxation, these closed strings may also contribute to the breakdown of the Stokes-Einstein relation in deeply supercooled liquids, in addition to dynamical heterogeneities [57].

Refer to caption
Figure 5: Energy dependence of (a) the fraction of string-like excitations and (b) the average string length, npsubscript𝑛𝑝n_{p}italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. (c) Scatter plots of the energy difference between the final and initial inherent structures (ISs) versus E𝐸Eitalic_E, for various npsubscript𝑛𝑝n_{p}italic_n start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. (d) Projection of displacement fields for representative excitations onto a 2D plane. We illustrate the emergence of a string as the system moves along its minimum energy path in Animation 1 [48].

Discussion – Through the novel ASEER algorithm, we have measured the density of excitations in a model glass-former up to the activation energy. Our findings confirm a shift in this density under cooling that can predict the fragility of the liquid. Most importantly, we have demonstrated that the geometry of excitations depends on both their energy and temperature, which in turn governs the stability of the overall medium. Excitations display a core whose properties scale with their energy and a far field component whose length scale is governed by temperature. These insights align with known observations on the geometry of relaxation in liquids, such as the increased predominance of strings under cooling.

These results support the idea that excitations are influenced by a dynamical transition. On one hand, the shift in the density of excitations mirrors the shift in the Hessian of the energy landscape predicted in high dimensions. This shift differs from a simple rescaling, as one might expect from a naive interpretation of elastic models where energies are scaled by a temperature-dependent elastic modulus. On the other hand, the excitation core follows scaling laws expected in the vicinity of a dynamical transition. According to this perspective, hop** processes suppress the divergence of the relaxation time at Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT but are very much influenced by the elastic instability associated with Tcsubscript𝑇𝑐T_{c}italic_T start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

Acknowledgments: We thank the Simons collaboration and L. Berthier, G. Biroli, C. Brito, C. Gavazzoni, E. Lerner, M. Muller, M. Popovic, M. Ozawa and A. Tahaei for discussions. M.P.C. discloses support for the research of this work from Singapore Ministry of Education [MOE-T2EP50221-0016]. M.W acknowledges support from the Simons Foundation Grant (No. 454953 Matthieu Wyart) and from the SNSF under Grant No. 200021-165509.

References

 

Supplementary Material for Density and geometry of excitations in supercooled liquids up to the activation energy

Numerical model, relaxation dynamics and activation energy

We consider a three-dimensional system of soft repulsive particles [43] with size σ𝜎\sigmaitalic_σ distributed as p(σ)σ3proportional-to𝑝𝜎superscript𝜎3p(\sigma)\propto\sigma^{-3}italic_p ( italic_σ ) ∝ italic_σ start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT in the range [σminsubscript𝜎min\sigma_{\rm min}italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT:2.2σmin2.2subscript𝜎min2.2\sigma_{\rm min}2.2 italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT]. This modern numerical model can be equilibrated up to experimentally comparable temperatures through the ‘swap’ algorithm [44, 45, 46, 47]. The pair interaction is given by

U(rij)=ϵ[(σijrij)10+l=03c2l(rijσij)2l]𝑈subscript𝑟𝑖𝑗italic-ϵdelimited-[]superscriptsubscript𝜎𝑖𝑗subscript𝑟𝑖𝑗10superscriptsubscript𝑙03subscript𝑐2𝑙superscriptsubscript𝑟𝑖𝑗subscript𝜎𝑖𝑗2𝑙U(r_{ij})=\epsilon\left[\left(\frac{\sigma_{ij}}{r_{ij}}\right)^{10}+\sum_{l=0% }^{3}c_{2l}\left(\frac{r_{ij}}{\sigma_{ij}}\right)^{2l}\right]italic_U ( italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) = italic_ϵ [ ( divide start_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT + ∑ start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 2 italic_l end_POSTSUBSCRIPT ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 italic_l end_POSTSUPERSCRIPT ] (4)

for rij<xc=1.4subscript𝑟𝑖𝑗subscript𝑥𝑐1.4r_{ij}\!<\!x_{c}\!=\!1.4italic_r start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT < italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1.4. We use a non-additive particle size σij=12(σi+σj)(10.1|σiσj|)subscript𝜎𝑖𝑗12subscript𝜎𝑖subscript𝜎𝑗10.1subscript𝜎𝑖subscript𝜎𝑗\sigma_{ij}\!=\!\frac{1}{2}(\sigma_{i}+\sigma_{j})(1\!-\!0.1|\sigma_{i}\!-\!% \sigma_{j}|)italic_σ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ( 1 - 0.1 | italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | ) to prevent crystallization and set c2lsubscript𝑐2𝑙c_{2l}italic_c start_POSTSUBSCRIPT 2 italic_l end_POSTSUBSCRIPT to enforce continuity at xcsubscript𝑥𝑐x_{c}italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT up to three derivatives. We studied systems of 𝒩=2000𝒩2000\mathcal{N}\!=\!2000caligraphic_N = 2000 particles of mass m𝑚mitalic_m at number density ρ=0.58𝜌0.58\rho=0.58italic_ρ = 0.58 in cubic simulation boxes with periodic boundary conditions. We express mass in units of m𝑚mitalic_m, temperature in units of ϵitalic-ϵ\epsilonitalic_ϵ, lengths in units ρ1/3superscript𝜌13\rho^{-1/3}italic_ρ start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT, and time in units of mσmin2/ϵ𝑚superscriptsubscript𝜎min2italic-ϵ\sqrt{m\sigma_{\rm min}^{2}/\epsilon}square-root start_ARG italic_m italic_σ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_ϵ end_ARG. We minimize the energy of configurations equilibrated at temperature T𝑇Titalic_T to produce ISs that we investigate with ASEER.

For this model system, in a previous work [40], we have (i) investigated the relaxation dynamics and shown that the self-scattering correlation function evaluated at the first peak of the static structure factor, a self-overlap function, and a total-overlap function, give consistent measures for the temperature dependence of the relaxation time τ𝜏\tauitalic_τ. All correlation functions satisfy the time-temperature superposition principle, which we exploit to measure the relaxation time at very low temperatures. (ii) estimated the microscopic time τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT influencing the relaxation time, τ=τ0eEa(T)/T𝜏subscript𝜏0superscript𝑒subscript𝐸𝑎𝑇𝑇\tau=\tau_{0}e^{E_{a}(T)/T}italic_τ = italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_T ) / italic_T end_POSTSUPERSCRIPT.

The evaluation of τ𝜏\tauitalic_τ and of τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT allows us to measure the activation energy regulating structural relaxation, Ea(T)=Tlog(τ/τ0)subscript𝐸𝑎𝑇𝑇𝜏subscript𝜏0E_{a}(T)=T\log(\tau/\tau_{0})italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_T ) = italic_T roman_log ( italic_τ / italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). Ea(T)subscript𝐸𝑎𝑇E_{a}(T)italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_T ) fixes the scale of the activation energy of the excitations that have a non-negligible probability of being activated during the relaxation dynamics. The ASEER algorithm crucially allows us to extract excitations with activation energy up and beyond Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT.

Consistency between SEER and ASEER

Refer to caption
Figure S1: We have shown in Fig. 2a that SEER and ASEER give consistent results for the density of excitations N(E)𝑁𝐸N(E)italic_N ( italic_E ) in the energy range they both access. The E𝐸Eitalic_E dependence of δ𝛿\deltaitalic_δ and |dr|2superscript𝑑𝑟2|dr|^{2}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT further confirms the consistency of these methods. We also illustrate median values (in cyan) at different E𝐸Eitalic_E up to Easubscript𝐸𝑎E_{a}italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, as shown in Fig. 3 in the main text. Data refer to T=0.4𝑇0.4T=0.4italic_T = 0.4.

Median and mean values satisfy the scaling relations

Refer to caption
Figure S2: In the main text, we have verified the scaling relations Eq. 3 by focusing on the median values of V=NPr𝑉𝑁subscript𝑃𝑟V\!=\!NP_{r}italic_V = italic_N italic_P start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, |dr|2superscript𝑑𝑟2|dr|^{2}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, δ𝛿\deltaitalic_δ, for the lowest energy excitations across 1000 samples. We show in these figures that the mean values behave analogously.

Robustness of Eq. 4 with respect to the definition of Eminsubscript𝐸minE_{\rm min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT

Refer to caption
Figure S3: We validated our proposed functional form (Eq. 4) for |dr|2(T,E)=|dr|2(Emin(T),E)superscript𝑑𝑟2𝑇𝐸superscript𝑑𝑟2subscript𝐸𝑇𝐸|dr|^{2}(T,E)=|dr|^{2}(E_{\min}(T),E)| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_T , italic_E ) = | italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_T ) , italic_E ) by adopting for Eminsubscript𝐸minE_{\rm min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT the median value of the lowest-energy excitations across 1000 independent samples. We prove the robustness of our prediction by showing that a data collapse is also obtained by replacing Eminsubscript𝐸minE_{\rm min}italic_E start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT with the median of the five lowest excitations E5subscript𝐸5E_{5}italic_E start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT (a), the median of the ten lowest excitations E10subscript𝐸10E_{10}italic_E start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT (b), or the gap energy Egsubscript𝐸𝑔E_{g}italic_E start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (c).

Architecture of high-energy excitations: the role of string-like excitations

Fig. 3 in the main text investigated the architecture of high-energy excitations that are not string-like. Including strings as well leave our results unchanged at small and intermediate energies. At very large energies, strings dominate and corrections are apparent for V𝑉Vitalic_V and δ𝛿\deltaitalic_δ, and less so |dr|2superscript𝑑𝑟2|dr|^{2}| italic_d italic_r | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Refer to caption
Figure S4: The analysis of the geometrical properties of non-string-like excitations illustrated in Fig. 3 in the main text is here repeated by also considering them.