This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Microscopic origin of liquid viscosity from unstable localized modes

Long-Zhou Huang1    Bingyu Cui2    Vinay Vaibhav3    Matteo Baggioli4 [email protected]    Yun-Jiang Wang1 [email protected] 1State Key Laboratory of Nonlinear Mechanics, Institute of Mechanics, Chinese Academy of Sciences, Beijing 100190, China
2School of Science and Engineering, The Chinese University of Hong Kong, Shenzhen, Guangdong, 518172, P. R. China
3Department of Physics “A. Pontremoli,” University of Milan, via Celoria 16, 20133 Milan, Italy
4Wilczek Quantum Center, School of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, China
Abstract

Viscosity is the resistance of a liquid to flow, governed by atomic-scale friction between its constituent atoms. While viscosity can be directly computed using the Green-Kubo formalism, this method does not fully elucidate the physical mechanisms underlying such momentum transport coefficient. In fact, the microscopic origin of liquid viscosity remains poorly understood. In this work, we calculate the viscosity of a Cu50Zr50\mathrm{Cu_{50}Zr_{50}} metallic liquid and a Kob-Andersen model in a large range of temperatures and compare the results with a theoretical formula based on nonaffine linear response and instantaneous normal mode theory. This analysis reveals that only unstable localized instantaneous normal modes (ULINMs) contribute to viscosity, providing a microscopic definition of viscosity as diffusive momentum transport facilitated by local structural excitations mediated by ULINMs as precursors. Leveraging on the concept of configurational entropy, a quantitative model to connect viscosity with ULINMs is also proposed and validated in both the Arrhenius and non-Arrhenius regimes. In summary, our work provides a microscopic definition of liquid viscosity and highlights the fundamental role of ULINMs as its building blocks, ultimately opening the path to an atomic-scale prediction of viscosity in liquids and glasses.

\color

blueIntroduction\colorblack – Viscosity is a fundamental transport coefficient in fluid dynamics [1] that reflects the ability of transferring momentum in the direction transverse to the motion of the liquid [2]. In the hydrodynamic regime, viscosity governs the flow of physical systems at very different energy scales, from honey [3] to electrons in ultra-clean metallic systems [4] and ultra-relativistic plasma of quarks and gluons [5].

From a macroscopic point of view, shear viscosity is defined by the linear response relation between shear stress and shear strain rate, that can be formalized using Green-Kubo formalism [6, 7]. This method allows for a direct estimate of the viscosity based on the time correlation function of shear stress, but it is limited to Newtonian fluids. Another established method to compute the viscosity is based on the Einstein-Stokes relation between the self-diffusion constant of a probe particle and the mobility of the fluid in which the latter is flowing [8, 9]. However, it is well-known that, as temperature approaches the glass transition, this relation breaks down hindering the applicability of this method.

Moreover, it is important to stress that both the Green-Kubo formula and the Einstein-Stokes relationship do not fully elucidate the microscopic physical mechanisms underlying momentum transport, and ultimately the viscosity itself. In particular, what is the direct link between microscopic dynamics and macroscopic viscosity in a fluid remains an open question. Using a hyperbole, the physical origin of liquid viscosity has not been fully disclosed yet.

In fact, this issue is part of a bigger problem that is the lack of an atomic-scale microscopic description of liquid dynamics and thermodynamics [10, 11], which is mostly due to the absence of a well-defined normal mode formalism for liquids, as achieved on the contrary for crystalline solids [12]. In this context of liquids, the existence of normal modes was already proposed by Zwanzig in 1967 [13], on the basis of Maxwell’s intuition [14] that liquids present solid characteristics at short time scales. These ideas were later formalized by Keyes and collaborators within the so-called instantaneous normal mode (INM) approach [15, 16], that is based on the diagonalization of the Hessian matrix in instantaneous liquid configurations.

Because of the absence of a well-defined equilibrium configuration in liquids, this procedure inevitably leads to the appearance of negative eigenvalues [17], or equivalently purely imaginary frequencies, that correspond to regions of the potential energy landscape with negative local curvature. Modes corresponding to positive eigenvalues are stable, while those with purely imaginary frequencies are called unstable INMs. The physical meaning of these unstable INMs remains matter of debate, contributing to the famous question of “what INMs are and are not[18].

Partially answering this question, in a series of works [19, 20], Keyes proposed that unstable INMs might provide the fundamental blocks to achieve a microscopic derivation of the liquid self-diffusion constant, highlighting their physical nature as relaxational processes mediated by jumps over potential barriers. It was later clarified that only unstable delocalized INMs contribute to the self-diffusion constant [21, 22] (see also [23, 24, 25, 26, 27, 23, 28] for related works and different definitions of this “diffusive” subset of INMs), providing a first connection between microscopic excitations and macroscopic transport in liquids, but leaving the physical meaning of the localized unstable modes unclear.

In this Letter, we investigate whether a microscopic definition for another fundamental transport coefficient in liquids – shear viscosity– can be achieved using a normal mode approach. In doing so, we will reveal the so-far unknown physical significance of unstable localized modes and their essential role for the microscopic origin of liquid viscosity.

We anticipate that a connection between high-frequency macroscopic elastic moduli and INMs in liquids was already proposed by Stratt [29], alluding to a possible connection between “certain” INMs and liquid viscosity as well. In this Letter, we corroborate that such a relation indeed exists and that unstable localized INMs are the building blocks of liquid viscosity. Our results provide a microscopic definition of viscosity, together with a parameter-free prediction for its value as a function of temperature.

\color

blueTheoretical model and methods\colorblack – According to the theoretical formalism presented in [30] based on nonaffine viscoelastic response [31], the shear viscosity derived from the imaginary part of the complex shear modulus is given by

η=3ρυ~(0)g(ω)Γ(ω)m2ω4𝑑ω,\eta=3\rho\,\tilde{\upsilon}(0)\int\frac{g\left(\omega\right)\Gamma\left(\omega\right)}{m^{2}\omega^{4}}\,d\omega, (1)

where ρ=N/V\rho=N/V represents the number density, NN is the number of atoms and VV is the volume of the system. Additionally, g(ω)g(\omega) is the density of states (DOS); mm is the atomic mass; Γ(ω)\Gamma(\omega) is the affine force field correlator [31] (see SM, Appendix D, for more details); υ~(0)\tilde{\upsilon}(0) is the zero-frequency limit of the spectral density, υ~(ω)=0υ(t)eiωt𝑑t\tilde{\upsilon}(\omega)=\int_{0}^{\infty}\upsilon(t)e^{-i\omega t}dt, with υ(t)\upsilon(t) the memory kernel arising from the coupling between tagged particles and thermal environment [32, 33]. All our following results and conclusions will rely on the validity of Eq. (1).

To derive the memory kernel, we use the projection operator formalism [34, 35, 36],

υ(t)=𝐅(ei(1𝒫)t𝐅)kBT=𝐅𝐅r(t)kBT=C¯𝐅𝐅(t)kBT,\upsilon(t)=\frac{\left\langle\mathbf{F}\left(e^{-i(1-\mathcal{P})\mathcal{L}t}\mathbf{F}\right)\right\rangle}{k_{\text{B}}T}=\frac{\langle\mathbf{FF}_{r}(t)\rangle}{k_{\text{B}}T}=\frac{\bar{C}_{\mathbf{FF}}(t)}{k_{\text{B}}T}, (2)

where 𝐅\mathbf{F} is the true force acting on the tagged particle. The random force 𝐅r=ei(1𝒫)t𝐅\mathbf{F}_{r}=e^{-i(1-\mathcal{P})\mathcal{L}t}\mathbf{F}, i-i\mathcal{L} is the Liouvillian operator corresponding to the unperturbed dynamics and 𝒫\mathcal{P} is the Mori projection operator along the direction of the velocity (see SM, Appendix F, for more details). Consequently, the zero-frequency limit of spectral density can be simply expressed as

υ~(0)=0C¯𝐅𝐅(t)kBT𝑑t.\tilde{\upsilon}(0)=\int\limits_{0}^{\infty}\frac{\bar{C}_{\mathbf{FF}}(t)}{k_{\text{B}}T}dt. (3)

Finally, we present two important remarks in applying Eq.(1) to liquid systems. First, the definition of the DOS in liquids is not straighforward. In our work, we will identify g(ω)g(\omega) with the INM density of states, allowing a direct connection to normal modes (that is not possible using the velocity auto-correlation function). Second, following this choice, it is necessary to define the integration range on ω\omega, that now takes values along both the real and imaginary axes. In the following, in order to find the correct range of integration, we will consider different choices and compare them directly with the simulation data. As already discussed, for the case of the self-diffusion constant [19], only unstable delocalized modes should be considered. It is therefore conceivable that a similar situation will hold for the shear viscosity.

In this study, two simulations systems are considered: (I) a Cu50Zr50\mathrm{Cu_{50}Zr_{50}} metallic liquid with Tg=690T_{\mathrm{g}}=690 K [37] and (II) a Kob-Andersen (KA) simulation model [38, 39]. Ten independent samples were obtained to reduce the fluctuation error of the simulations. More details about the models and simulations are provided in the Supplementary Material (SM). The results in the main text refer to the metallic liquid; analogous results for the KA model are presented in the SM and present the same features showing the universality of microscopic origin of viscosity.

We notice that, once the range of integration on ω\omega in Eq.(1) is fixed, our procedure will not involve any free fitting parameter. Therefore, our results could be taken as a direct test of the validity of Eq.(1) that, to the best of our knowledge, has never been verified.

Refer to caption
Figure 1: (a) INM DOS g(ω)g(\omega) and (b) affine force field correlator Γ(ω)/m2\Gamma(\omega)/m^{2} versus frequency at different temperature. (c) Auto-correlation function of the random force C¯𝐅𝐅(t)\bar{C}_{\mathbf{FF}}(t) versus time at different temperature. (d) Temperature dependence of the zero-frequency limit of spectral density υ~(0)\tilde{\upsilon}(0). Solid black circles are calculated using Eq. (5), while hollow circles by the projection operator technique, Eq. (3).
\color

blueMicroscopic description of liquid dynamics\colorblack – We performed INM analysis of the Cu50Zr50\mathrm{Cu_{50}Zr_{50}} metallic liquid for temperatures between 800800 K and 22002200 K. The corresponding INM DOS is shown in Fig. 1(a), following the common practice of representing imaginary frequencies along the negative axes [15]. The arrows indicate the directions along which TT increases. We notice that g(ω)g(\omega) is linear at low frequency (see SM, Fig. S3, for more details), in agreement with theoretical expectations (e.g., [19, 40, 41]) and experimental observations [42, 43].

Panels (b) and (c) of Fig. 1 display respectively the re-scaled affine force field versus frequency and the auto-correlation function of the random force versus time at different temperatures. From the memory kernel defined in Eq. (2), we introduce the friction coefficient

γ=0υ(t)𝑑t.\gamma=\int\limits_{0}^{\infty}\upsilon(t)dt. (4)

Using the Einstein’s relation γ=kBT/D\gamma=k_{\mathrm{B}}T/{D}, where DD is the long-range diffusion constant of the tagged particles, and comparing Eqs. (3) and (4), we get

υ~(0)kBTD.\tilde{\upsilon}(0)\cong\frac{k_{\text{B}}T}{D}. (5)

We find that computing υ~(0)\tilde{\upsilon}(0) in this way gives results compatible with those obtained using the projection operator technique, Eq.(3), as shown explicitly in Fig. 1(d). The data also follow the empirical Vogel–Fulcher–Tammann (VFT) equation [44, 45, 46], η(T)=η0exp(BVFTTT0)\eta(T)=\eta_{0}\exp\left(\frac{B_{\mathrm{VFT}}}{T-T_{0}}\right), shown with a red line in the same panel.

As anticipated, INMs can be classified into stable and unstable modes depending on the sign of the corresponding eigenvalues. The latter can be further divided into localized and delocalized, depending whether the number of particles participating in the mode is intensive or extensive. Bembenek and Laird originally argued that only unstable delocalized modes are related to diffusion [21, 47], while unstable localized modes do not contribute to it. However, there are indications that unstable localized modes may play a fundamental role for other dynamical properties, such as structural relaxation [48, 49].

In order to quantify the degree of localization, we consider the participation ratio

Pω=[Ni=1N(𝐞iω𝐞iω)2]1,P_{\omega}=\left[N\sum_{i=1}^{N}(\mathbf{e}_{i}^{\omega}\cdot\mathbf{e}_{i}^{\omega})^{2}\right]^{-1}, (6)

where 𝐞iω\mathbf{e}_{i}^{\omega} is the contribution of particle ii to the normalized eigenvector corresponding to eigenfrequency ω\omega. PωP_{\omega} takes values between 1/N1/N and 11, with the former corresponding to an extremely localized mode with only one particle participating, while the latter to a fully delocalized mode involving all particles (see SM, Fig. S11, for direct visualization of these two types of modes and Ref. [50] for details about PωP_{\omega}).

In order to identify the mobility edge separating localized and delocalized modes, we follow [21, 47] and perform a finite size scaling analysis and compute PωP_{\omega} for different NN. In Fig. 2(a), we show PωP_{\omega} as a function of frequency for the unstable branch of INMs and different system sizes. From there, we can observe that PωP_{\omega} becomes system dependent below a certain threshold frequency, that for the CuZr\mathrm{CuZr} glass-forming system corresponds to Pω0.37P_{\omega}\approx 0.37. Intensive modes, with PωP_{\omega} below the mobility edge, are localized, while extensive modes with Pω0.37P_{\omega}\gtrapprox 0.37 are delocalized.

The participation ratio as a function of frequency for different temperatures is shown in Fig. 2(b). By repeating the same finite size scaling analysis for different TT, we found that the value Pω0.37P_{\omega}\approx 0.37, defining the mobility edge, is in first approximation independent of temperature (see SM, Fig. S10, for a direct proof of this statement). By then searching for the corresponding (imaginary) frequency (see horizontal dashed line in Fig. 2(b)), we have derived the mobility edge frequency that is plotted in Fig. 2(c) as a function of temperature. We find that the mobility edge frequency increases (in absolute value) by increasing temperature, consistent with the idea that unstable modes tend to be more localized towards the glass transition temperature [21].

Refer to caption
Figure 2: (a) Variation of the participation ratio as a function of (imaginary frequency) at 15001500 K upon changing the size of the simulation box. (b) Participation ratio versus frequency at different temperatures. The dotted horizontal line represents the mobility edge threshold Pω=0.37P_{\omega}=0.37. (c) Temperature dependence of the mobility edge frequency. (d) Fraction of stable, unstable, unstable localized and unstable delocalized modes as a function of temperature.

Finally, in Fig. 2(d), we respectively show the fraction of stable fsf_{\mathrm{s}}, unstable fuf_{\mathrm{u}}, unstable localized fuLf_{\mathrm{u}}^{\mathrm{L}} and unstable delocalized fuDf_{\mathrm{u}}^{\mathrm{D}} INMs in the temperature range considered. We find that the fraction of stable modes slowly decreases by increasing temperature, while all the other fractions increases with temperature following a similar trend. Once the temperature reaches the glass transition, unstable delocalized modes disappear and all unstable modes become localized, consistent with previous results [22].

We have now independently derived all the physical quantities entering in Eq.(1) and analyzed in detail the different types of modes over which the integration therein could be in principle performed. Before moving to the computation of the liquid viscosity, we notice that, due to the limited size of the simulation box LL, the integration in Eq.(1) will be always performed up to a lower cutoff frequency |ωmin|=(2πvs)/L|\omega_{\min}|=(2\pi v_{s})/L, where the speed of sound vsv_{s} is given by B/ρ¯\sqrt{B/\bar{\rho}}, with BB the bulk modulus and ρ¯\bar{\rho} the mass density (see SM, Appendix H, for more details).

\color

blueLiquid viscosity\colorblack – Traditionally, the viscosity can be computed using the Green-Kubo formula [6, 7],

η=VkBT0σαβ(t)σαβ(0)𝑑t,\eta=\frac{V}{k_{\mathrm{B}}T}\int\limits_{0}^{\infty}\left\langle\sigma_{\alpha\beta}(t)\sigma_{\alpha\beta}(0)\right\rangle dt, (7)

where σαβ\sigma_{\alpha\beta} is the α,β\alpha,\beta component of the stress tensor (with α,β=x,y,x\alpha,\beta=x,y,x). In Eq.(7), \langle\cdots\rangle indicates ensemble average. Alternatively, the viscosity can be obtained from non-equilibrium molecular dynamics using the linear response relation η=σαβ/ε˙\eta=\sigma_{\alpha\beta}/\dot{\varepsilon}, where ε˙\dot{\varepsilon} is the shear strain rate. We will indicate this second route as shear MD.

Refer to caption
Figure 3: (a) Comparison of viscosities from the Green-Kubo formula Eq.(7) (black circles), MD non-equilibrium shear (red crosses), the present calculation based on Eq.(1) and unstable localized INMs (blue symbols), and experimental data from Refs. (Markus et al. [51]) (purple diamonds) and (Mauro et al. [52]) (green crosses). The solid line represents the VFT fit. (b) Comparison of viscosities from Green-Kubo formula (black circles) and different INMs including unstable localized (blue symbols), unstable delocalized (red inverted triangles), stable (purple triangles) and all modes (green diamonds).

In Fig. 3(a), we compare the viscosity calculated by different methods: black symbols corresponds to the Green-Kubo formula Eq.(7), red crosses to shear MD and blue symbols are the results obtained from Eq.(1), without free parameters, where the integration is performed over only the unstable localized INMs (ULINMs). For comparison, we also show the experimental data from Refs. [51]) (purple diamonds) and [52] (green crosses). At high temperature (above T1000T\approx 1000 K, 1.51.5 times the glass transition temperature), Green-Kubo formula and shear MD are in good agreement. However, when temperature drops, the shear viscosity becomes strain rate dependent and the linear response formula (hence, the shear MD method) is not anymore applicable, giving unreliable predictions for η\eta.

Most importantly, we find that the viscosity computed via the Green-Kubo formula is in perfect agreement with the results of Eq.(1) using only ULINMs, and both data are well described by the VFT equation. This is the main result of this work. Same results are obtained for the Kob-Anderson liquid model and are presented in the SM, Fig. S1, confirming the excellent agreement found for the CuZr metallic system and the universality of our conclusions.

For completeness, in Fig. 3(b) we show a comparison between the viscosity obtained from Green-Kubo and the viscosity obtained from the normal mode formula Eq.(1) using different integration regimes or, in other words, different types of INMs as the microscopic building blocks for viscosity. We find that only the estimate using ULINMs is in agreement with the Green-Kubo data. All other choices give results that are almost an order of magnitude away from the correct value of η\eta, especially at high temperature. This suggests that unstable localized modes are the microscopic origin of liquid viscosity, and that other excitations are unrelated to this macroscopic transport property.

Inspired by Adam-Gibbs theory [53], and motivated by recent results on the relation between liquid diffusion and excess entropy [54], we construct a theoretical model to quantitatively correlate ULINMs with macroscopic viscosity. Following the ideas in [22], we make a similar attempt to link ULINMs to configurational entropy, and then viscosity. First, in Fig. 4(a) we find that the configurational entropy SconfS_{\mathrm{conf}} displays a linear relationship with ln(fuL)\ln(f_{\mathrm{u}}^{\mathrm{L}}). Here, the configurational entropy is estimated by subtracting the vibrational entropy from the total entropy obtained by thermodynamic integration, as detailed in [55]. Second, we observe that the viscosity follows a direct non-Arrhenius relation with the configurational entropy, ηAexp(B/(TSconf )λ)\eta\propto A\exp\left(B/\left(TS_{\text{conf }}\right)^{\lambda}\right), with λ=2.10\lambda=2.10. We notice the analogy with the connection between the α\alpha relaxation time and Sconf S_{\text{conf }}, usually rationalized using random first-order transition theory [53, 56, 57]. By combining these two observations, we conclude that ηAexp(B/(T[aln(fuL)+b])λ)\eta\propto A\exp\left(B/\left(T\left[a\ln\left(f_{\mathrm{u}}^{\mathrm{L}}\right)+b\right]\right)^{\lambda}\right), as demonstrated using the simulation data in Fig. 4(c). This result provides a further link between viscosity and ULINMs and validate their intimate relation, providing a bridge between microscopic and macroscopic physics in liquids.

Refer to caption
Figure 4: (a) Linear scaling between SconfS_{\mathrm{conf}} and ln(fuL)\ln(f_{\mathrm{u}}^{\mathrm{L}}). (b) Viscosity scales as [TSconf]λ[TS_{\mathrm{conf}}]^{-\lambda}, with λ=2.10\lambda=2.10. (c) Viscosity versus [T(aln(fuL)+b)]λ[T(a\ln(f_{\mathrm{u}}^{\mathrm{L}})+b)]^{-\lambda}. The solid line indicates the theoretical fit with a=0.92,b=4.15a=0.92,\ b=4.15 and λ=2.10\lambda=2.10.
\color

blueConclusions\colorblack – In this Letter, we have performed MD simulations combined with instantaneous normal mode analysis to disclose the microscopic origin of viscosity in liquids. Our results indicate that unstable localized INMs are the microscopic excitations responsible for this process. Moreover, our work provides a parameter-free formula to predict from normal modes the viscosity of liquids and its temperature dependence, defying the famous Landau argument that is “impossible to derive any general formulae giving a quantitative description of the properties of a liquid[58].

More in general, our results contribute to answering the question of what unstable INMs are and are not. Unstable delocalized modes are the building blocks for liquid self-diffusion, while the complementary ULINMs are the microscopic facilitators underlying liquid viscosity.

Interestingly, this answer complements several previous analyses. In particular, in Ref.[49] it was shown that localized soft modes are the origin of irreversible structural reorganization. Here, we give a microscopic definition of these localized soft modes and we propose that they should be identified with ULINMs. This is also consistent with the results of [48] that showed that ULINMs act as facilitators of the liquid dynamics, located at the boundary regions between mobile and immobile clusters generated by dynamical heterogeneity. Moreover, our findings are aligned with the idea that, in glass-forming liquid, excitations are localized and relaxation is hierarchical [59]. Finally, viscosity being governed by localized modes is also reflected in the recent results of [60], where it was found that the change in the local barrier energy caused by the fundamental rearrangement or excitation of a few particles (hence, localized modes) determines the change in the activation energy and the dynamics of the supercooled liquids.

In summary, it was repeatedly recognized that “the viscosity of liquids is a subject which has hitherto been without any general theoretical basis[61]. By identifying the microscopic carriers of momentum transport, our work suggests a possible microscopic and quantitative theory of liquid viscosity, complementing the previous approaches [1] and providing a new path towards a complete understanding of liquid dynamics.

\color

blueAcknowledgments \colorblack – We thank A. Zaccone for collaboration at an early stage of this work and for valuable comments and suggestions. We would like to thank J. Douglas, M. Pica Ciamarra, J. Dyre, T. Keyes, J. Moon and A. Lemaitre for useful comments and discussion. This work was financially supported by the Strategic Priority Research Program (grants nos. XDB0620103 and XDB0510301) and the Youth Innovation Promotion Association of Chinese Academy of Sciences, and the National Natural Science Foundation of China (grant no. 12072344). The numerical calculations in this study were carried out on the ORISE Super-computer. M.B. acknowledges the support of the Shanghai Municipal Science and Technology Major Project (Grant No.2019SHZDZX01) and the sponsorship from the Yangyang Development Fund.

References

Supplementary Material

Microscopic origin of liquid viscosity from unstable localized modes
Long-Zhou Huang, Bingyu Cui, Vinay Vaibhav,
Matteo Baggioli, Yun-Jiang Wang

In this Supplementary Material (SM), we provide further details about the numerical and analytical computations presented in the main text. Moreover, we show further analysis on a Kob-Andersen model to corroborate our findings.

Appendix A Metallic liquid

A semi-empirical potential function is used to simulate the interactions in a Cu50Zr50\mathrm{Cu_{50}Zr_{50}} metallic glass with a simulation box of 4000 atoms [37]. The liquid and glass samples are prepared with the standard heating-cooling method. First, a crystalline CuZr\mathrm{CuZr} phase was heated from 300 K to 2300 K for melting. And then the liquid is thermally equilibrated for 1 ns at 2300 K which is much longer than the α\alpha relaxation time. Then the equilibrium liquid was quenched to supercooled liquid to 300 K with constant cooling rate of 101010^{10} K/s. The above sample preparation process was carried out under the NPT ensemble with periodic boundary condition. The Nose-Hoover thermostat was used to control temperature and the barostat was used to keep pressure to zero [62]. The kinetic properties were estimated using standard microcanonical (NVE) ensemble after sufficient equilibration at each specific temperature. For statistical purpose, ten independent samples were simulated to reduce the fluctuation of physical quantities. The velocity Verlet algorithm was used to integrate Newton’s equations of motion with a time step of 1 fs.

Appendix B Kob-Andersen model

The Kob-Andersen model is binary Lennard-Jones mixture introduced by Kob and Andersen [38, 39] and defined by the following LJ potential,

Uab(r)=4Yab[(Xabr)12(Xabr)6],U_{ab}(r)=4Y_{ab}\left[\left(\frac{X_{ab}}{r}\right)^{12}-\left(\frac{X_{ab}}{r}\right)^{6}\right], (S1)

where a,b{A,B}a,b\in\left\{A,B\right\}; here YAA=1.0,XAA=1.0,YAB=1.5,XAB=0.8,YBB=0.5Y_{AA}=1.0,X_{AA}=1.0,Y_{AB}=1.5,X_{AB}=0.8,Y_{BB}=0.5, and XBB=0.88X_{BB}=0.88. The KA model consists of particles A and B with the same mass mm in a ratio of 80:20. We use reduced units, where length is in the unit of XAAX_{AA}, temperature YAA/kBY_{AA}/k_{\mathrm{B}}, and time (mXAA2/YAA)1/2(mX_{AA}^{2}/Y_{AA})^{1/2}. Under periodic boundary conditions, the system contains 4000 particles. Under the NVT ensemble, the fixed density is 1.185 and the cooling rate is 3.3×1063.3\times 10^{-6}. The kinetic properties were calculated using standard microcanonical (NVE) ensemble. Ten independent samples were obtained to reduce the fluctuation error of the simulations. The time step is 0.005. For the KA model, the temperature TT will be always displayed in reduced LJ units.

Refer to caption
Figure S1: (a) Comparison of the viscosity of Kob-Anderson model by Green-Kubo approach and Eq. (1) in the main text based on unstable localized modes. (b) Viscosity versus the fraction of unstable localized modes.

As it can be seen from the Fig. S1(a), the relationship between viscosity and unstable localized mode of Kob-Anderson model is similar to that found in CuZr\mathrm{CuZr} metallic system, which justifies that the generality of our observation that unstable localized mode are is the physical origin of viscosity in liquid. As shown in Fig. S1(b), the viscosity decreases by increasing the fraction of unstable localized modes. This is consistent with the fact that fuLf_{\mathrm{u}}^{\mathrm{L}} grows with temperature and the viscosity decreases while increasing temperature in the liquid phase.

Appendix C Instantaneous normal modes

The dynamical matrix for the instantaneous liquid configuration at a given temperature (instantaneous Hessian matrix) is given by

𝐃=2Uri,αrj,βmimj,\mathbf{D}=\frac{\partial^{2}U}{\partial r_{i,\alpha}\partial r_{j,\beta}\sqrt{m_{i}m_{j}}}, (S2)

where mim_{i} is the mass of particle ii and ri,αr_{i,\alpha} is the coordinate (x,yx,y or zz) of particle ii. We directly diagonalize the dynamical matrix and calculate the vibrational density of states as

g(ω)=limΔω0ΔnΔω=13N3pδ(ωωp),g(\omega)=\lim_{\Delta\omega\to 0}\frac{\Delta n}{\Delta\omega}=\frac{1}{3N-3}\sum_{p}\delta(\omega-\omega_{p}), (S3)

where ωp\omega_{p} is the eigenfrequency. Ten independent samples were obtained to reduce the fluctuation error of the simulations.

The results for the INM DOS are shown in Fig. S2 and present similar features as for the metallic liquid discussed in the main text.

Refer to caption
Figure S2: Instantaneous normal mode DOS for the Kob-Anderson model at different temperature. The arrows indicate the trend of change with increasing temperature.

A zoom of the low frequency region of g(ω)g(\omega) for the CuZr\mathrm{CuZr} metallic glass-forming liquid is shown in Fig. S3. The data confirm the linear scaling g(ω)=a(T)|ω|g(\omega)=a(T)|\omega| (with a(T)a(T) growing with temperature) discussed in the main text.

Refer to caption
Figure S3: Zoom of the low frequency region of g(ω)g(\omega) for the CuZr\mathrm{CuZr} metallic glass-forming liquid. A linear behavior g(ω)=a(T)|ω|g(\omega)=a(T)|\omega| emerges at low frequency. From our simulation data, it is also evident that a(T)a(T) grows with TT.

Appendix D Affine force field correlator

According to nonaffine lattice dynamics theory [31], the affine force field is obtained from

𝚵i,xy=𝐟iεxy|ε0=𝐟i1𝐟i0εxy,\mathbf{\Xi}_{i,xy}=\frac{\partial\mathbf{f}_{i}}{\partial\varepsilon_{xy}}\Bigg{|}_{\varepsilon\to 0}=\frac{\mathbf{f}_{i}^{1}-\mathbf{f}_{i}^{0}}{\varepsilon_{xy}}, (S4)

where 𝐟i1\mathbf{f}_{i}^{1} is the force on the ii atom in deformed state; 𝐟i0\mathbf{f}_{i}^{0} is the force on the ii atom in undeformed state; εxy\varepsilon_{xy} is the simple shear strain. The projection of the eigenvectors 𝐯ωp\mathbf{v}^{\omega_{p}} of the Hessian matrix onto the affine force field are then given by

Ξ^p,xy=𝐯ωp𝚵xy,\hat{\Xi}_{p,xy}=\mathbf{v}^{\omega_{p}}\cdot\mathbf{\Xi}_{\mathrm{xy}}, (S5)

and the affine force field correlator is derived from

Γ(ω)=Ξ^p,xy2ωpϵ[ω,ω+dω].\Gamma(\omega)=\left\langle\hat{\Xi}_{p,xy}^{2}\right\rangle_{\omega_{p}\epsilon[\omega,\omega+d\omega]}. (S6)

A shear strain of 0.0001 is applied in our simulation of Kob-Anderson model and CuZr\mathrm{CuZr} metallic glass. The affine force field correlator of the Kob-Anderson model is shown as a function of frequency, for the liquid sample at different temperature as shown in the Fig. S4.

Refer to caption
Figure S4: Affine force field correlator versus frequency at different temperature for the Kob-Anderson model. All the curves are obtained by averaging over 10 independent samples.

Appendix E Mean square displacement

The diffusion constant can be obtained from the mean square displacement (MSD) as a function of time

Dα=12limτ[ri,α(t+τ)ri,α(t)]2τ.D_{\alpha}=\frac{1}{2}\lim_{\tau\rightarrow\infty}\frac{\left\langle\left[r_{i,\alpha}(t+\tau)-r_{i,\alpha}(t)\right]^{2}\right\rangle}{\tau}. (S7)

The average diffusion constant in an isotropic system in 3D is given by

D=13α=13Dα.D=\frac{1}{3}\sum_{\alpha=1}^{3}D_{\alpha}. (S8)
Refer to caption
Figure S5: The mean square displacement as a function of time. (a) CuZr\mathrm{CuZr} metallic liquid and (b) Kob-Anderson model.

Appendix F Projection operator technique

According to the projection operator technique [34, 35, 36], the memory kernel

υ(t)=𝐅(ei(1𝒫)t𝐅)kBT=𝐅𝐅r(t)kBT=C¯𝐅𝐅(t)kBT,\upsilon(t)=\frac{\left\langle\mathbf{F}\left(e^{-i(1-\mathcal{P})\mathcal{L}t}\mathbf{F}\right)\right\rangle}{k_{\text{B}}T}=\frac{\langle\mathbf{FF}_{r}(t)\rangle}{k_{\text{B}}T}=\frac{\bar{C}_{\mathbf{FF}}(t)}{k_{\text{B}}T}, (S9)

where 𝐅\mathbf{F} the true force acting on the tagged particle. Based on the second-order backward algorithm, C¯𝐀𝐀(t)=𝐀0𝐀t\bar{C}_{\mathbf{AA}}(t)=\left\langle\mathbf{A}_{0}\mathbf{A}_{t}^{-}\right\rangle, where 𝐀\mathbf{A} is an observable, and 𝐀t=ei(1𝒫)t𝐀0\mathbf{A}_{t}^{-}=e^{-i(1-\mathcal{P})\mathcal{L}t}\mathbf{A}_{0}. 𝐅t\mathbf{F}_{t}^{-} evolves according to the backward orthogonal dynamics

𝐅t+δt(𝐪,𝐩)=𝐅t(𝐪δt,𝐩δt)+0δt𝐅u(𝐪δt,𝐩δt)𝐏0𝐅tu𝐏02𝑑u,\mathbf{F}_{t+\delta t}^{-}(\mathbf{q},\mathbf{p})=\mathbf{F}_{-t}^{-}\left(\mathbf{q}^{\delta t},\mathbf{p}^{\delta t}\right)+\int_{0}^{\delta t}\mathbf{F}_{-u}\left(\mathbf{q}^{\delta t},\mathbf{p}^{\delta t}\right)\frac{\left\langle\mathbf{P}_{0}\mathbf{F}_{t-u}^{-}\right\rangle}{\left\langle\mathbf{P}_{0}^{2}\right\rangle}du, (S10)

where 𝐪\mathbf{q} and 𝐩\mathbf{p} are respectively the atomic positions and momenta. The second-order propagator

Fn+1(m)=Fn(m)+β(n)Fn(m)δt2+Fn(m)1δt2κ(n)[γ(n)+Δ(n)β(n)δt2]δt2,F_{n+1}^{-}(m)=F_{n}^{-}(m)+\beta(n)F_{n}(m)\frac{\delta t}{2}+\frac{F_{n}(m)}{1-\frac{\delta t}{2}\kappa(n)}\biggl{[}\gamma(n)+\Delta(n)\beta(n)\frac{\delta t}{2}\biggr{]}\frac{\delta t}{2}, (S11)

with

β(n)=m=1NtrajNcorrPn(m)Fn(m)m=1NtrajNcorrPn(m)2\beta(n)=\frac{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)F_{n}^{-}(m)}{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)^{2}} (S12)
κ(n)=m=1NtrajNcorrFn(m)Pn(m)m=1NtrajNcorrPn(m)2\kappa(n)=\frac{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}F_{n}(m)P_{n}(m)}{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)^{2}} (S13)
γ(n)=m=1NtrajNcorrPn(m)Fn(m+1)m=1NtrajNcorrPn(m)2\gamma(n)=\frac{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)F_{n}^{-}(m+1)}{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)^{2}} (S14)
Δ(n)=m=1NtrajNcorrFn(m)Pn(m+1)m=1NtrajNcorrPn(m)2.\Delta(n)=\frac{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}F_{n}(m)P_{n}(m+1)}{\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}P_{n}(m)^{2}}. (S15)

As a result, based on the trajectories of the atoms in each configuration we can get 𝐅t\mathbf{F}_{t}^{-}. Finally, C¯FF\bar{C}_{FF} can be obtained,

C¯FF(nδt)=1NtrajNcorrm=1NtrajNcorrFn(m)Fn(m).\bar{C}_{FF}(n\delta t)=\frac{1}{N_{\mathrm{traj}}-N_{\mathrm{corr}}}\sum_{m=1}^{N_{\mathrm{traj}}-N_{\mathrm{corr}}}F_{n}(m)F_{n}^{-}(m). (S16)
Refer to caption
Figure S6: Auto-correlation function versus time at different temperatures T[0.6,2]T\in[0.6,2] for the KA model. LJ units are used for the temperature TT.

As shown explicitly in Fig. S7, the results for the friction coefficient obtained using 0C¯𝐅𝐅(t)/kBT𝑑t\int_{0}^{\infty}\bar{C}_{\mathbf{FF}}(t)/k_{\mathrm{B}}Tdt and kBT/Dk_{\mathrm{B}}T/D are in good agreement. Therefore, we will use kBT/Dk_{\mathrm{B}}T/D, instead of 0C¯𝐅𝐅(t)/kBT𝑑t\int_{0}^{\infty}\bar{C}_{\mathbf{FF}}(t)/k_{\mathrm{B}}Tdt, to calculate υ~(0)\tilde{\upsilon}(0) at low temperatures.

Refer to caption
Figure S7: Consistency of 0C¯𝐅𝐅(t)/kBT𝑑t\int_{0}^{\infty}\bar{C}_{\mathbf{FF}}(t)/k_{\mathrm{B}}Tdt and kBT/Dk_{\mathrm{B}}T/D for calculating the friction coefficient. (a) CuZr\mathrm{CuZr} glass-forming liquid and (b) Kob-Anderson model.

Appendix G Participation ratio

The participation ratio of Kob-Anderson model is calculated and analyzed in Fig. S8 and displays features similar to the case of the CuZr\mathrm{CuZr} metallic glass-forming liquid discussed in the main text.

Refer to caption
Figure S8: The participation ratio of Kob-Anderson model. (a) Variation of participation ratio with mode size at temperature 1; (b) Participation ratio versus frequency at different temperature. The dotted horizontal line represents Pω=0.27P_{\omega}=0.27.

As done in the main text, we can calculate the dependence of the mobility edge frequency on temperature for the KA model, as shown in the Fig. S9.

Refer to caption
Figure S9: Temperature dependence of the mobility edge frequency for the Kob-Anderson model.

Meanwhile, variation of the participation ratio of CuZr\mathrm{CuZr} metallic glass-forming liquid as a function of (imaginary frequency) at 18001800 K and 21002100 K upon changing the size of the simulation box as shown in the Fig. S10. A typical localized and delocalized mode are shown by the magnitude of each particles vibrational vector as shown in the Fig. S11. Several high amplitude eigenvectors at local frequencies indicate strong localization, indicating that the energy associated with these modes is confined to a small spatial region.

Refer to caption
Figure S10: Variation of the participation ratio of CuZr\mathrm{CuZr} metallic glass-forming liquid as a function of (imaginary frequency) at (a) 18001800 K and (b) 21002100 K upon changing the size of the simulation box.
Refer to caption
Figure S11: Spatial projection of the eigenvectors on the xy-plane of each particle corresponding to (a) unstable localized mode (ω=25.91THz\omega=-25.91\ \mathrm{THz}), (b) unstable delocalized mode (ω=1.73THz\omega=-1.73\ \mathrm{THz}), (c) stable delocalized mode (ω=1.09THz\omega=1.09\ \mathrm{THz}) and (d) stable localized mode (ω=50.04THz\omega=50.04\ \mathrm{THz}) of CuZr\mathrm{CuZr} metallic liquid at 1500 K.

Appendix H Bulk modulus

The elastic constants are calculated using LAMMPS [63], and then bulk modulus can be obtained according to the Voigt-Reuss-Hill approximation [64]. In the Fig. S12, it is shown that the bulk modulus of CuZr\mathrm{CuZr} decreases with increasing temperature. Fig. S13 shows the value of the cutoff frequency, |ωmin|=(2πB/ρ¯)/L|\omega_{\min}|=(2\pi\sqrt{B/\bar{\rho}})/L, as a function of temperature for the CuZr\mathrm{CuZr} metallic liquid.

Refer to caption
Figure S12: Temperature dependence of bulk modulus for the CuZr\mathrm{CuZr} metallic glass-forming system.
Refer to caption
Figure S13: Temperature dependence of cutoff frequency ωmin\omega_{\min} for the CuZr\mathrm{CuZr} metallic glass-forming system.

Appendix I Vogel-Fulcher-Tammann (VFT) equation

The VFT equation reads [44, 45, 46]

η(T)=η0exp(BVFTTT0),\eta(T)=\eta_{0}\exp\left(\frac{B_{\mathrm{VFT}}}{T-T_{0}}\right), (S17)

where η0\eta_{0} is the viscosity in the infinite temperature limit, T0T_{0} is the temperature at which the viscosity diverges and BVFTB_{\mathrm{VFT}} denotes the fragility strength coefficient related with the activation barrier.

Appendix J Generalized Langevin equation (GLE)

The linear generalized Langevin equation (GLE) is given by [65]

𝐩˙(t)=mω¯2𝐪(t)0tυ(tτ)𝐯(τ)𝑑τ+𝐅r(t),\dot{\mathbf{p}}(t)=-m\bar{\omega}^{2}\mathbf{q}(t)-\int_{0}^{t}\upsilon(t-\tau)\mathbf{v}(\tau)d\tau+\mathbf{F}_{r}(t), (S18)

where ω¯\bar{\omega} is the frequency of an effective harmonic oscillator. Multiplying the linear GLE with 𝐩(0)\mathbf{p}(0) and taking a canonical ensemble average, we can get

C˙𝐩𝐩(t)=mω¯2C𝐩𝐪(t)0tυ(tτ)C𝐩𝐯(τ)𝑑τ.\dot{C}_{\mathbf{pp}}(t)=-m\bar{\omega}^{2}C_{\mathbf{pq}}(t)-\int_{0}^{t}\upsilon(t-\tau)C_{\mathbf{pv}}(\tau)d\tau. (S19)

By using the relation

C𝐩𝐪(t)=0tC𝐩𝐯(τ)𝑑τ,C_{\mathbf{pq}}(t)=\int_{0}^{t}C_{\mathbf{pv}}(\tau)d\tau, (S20)

Eq. S19 can be simplified in the following form

C˙𝐩𝐩(t)=0tK(tτ)C𝐩𝐯(τ)𝑑τ,\dot{C}_{\mathbf{pp}}(t)=-\int_{0}^{t}K(t-\tau)C_{\mathbf{pv}}(\tau)d\tau, (S21)

where

K(t)=υ(t)+mω¯2.K(t)=\upsilon(t)+m\bar{\omega}^{2}. (S22)

Upon Fourier transforming Eqs. (S21) and (S22), one obtains

K~(ω)=C𝐩𝐩(0)C~𝐩𝐯(ω)iω,\tilde{K}(\omega)=\frac{C_{\mathbf{pp}}(0)}{\tilde{C}_{\mathbf{pv}}(\omega)}-i\omega, (S23)
K~(ω)=υ~(ω)+πmw¯2δ(ω)imw¯2ω.\tilde{K}(\omega)=\tilde{\upsilon}(\omega)+\pi m\bar{w}^{2}\delta(\omega)-i\frac{m\bar{w}^{2}}{\omega}. (S24)

Comparing Eqs. (S23) and (S24), we finally obtain

υ~(0)=limω0υ~(ω)limω0K~(ω)=limω0C𝐩𝐩(0)C~𝐩𝐯(ω)=C𝐩𝐩(0)C~𝐩𝐯(ω)=mkBTmC~𝐯𝐯(0)=kBT0𝐯(𝐭)𝐯(𝟎)𝑑t=kBTD.\tilde{\upsilon}(0)=\lim_{\omega\to 0}\tilde{\upsilon}(\omega)\approx\lim_{\omega\to 0}\tilde{K}(\omega)=\lim_{\omega\to 0}\frac{C_{{{}_{\mathbf{pp}}}}(0)}{\tilde{C}_{{{}_{\mathbf{pv}}}}(\omega)}=\frac{C_{{{}_{\mathbf{pp}}}}(0)}{\tilde{C}_{{{}_{\mathbf{pv}}}}(\omega)}=\frac{mk_{{{}_{\mathrm{B}}}}T}{m\tilde{C}_{{{}_{\mathbf{vv}}}}(0)}=\frac{k_{{{}_{\mathrm{B}}}}T}{\int_{0}^{\infty}\left\langle\mathbf{v(t)v(0)}\right\rangle dt}=\frac{k_{{{}_{\mathrm{B}}}}T}{D}. (S25)