View metadata, citation and similar papers at core.ac.uk brought to you by CORE

provided by Queen Mary Research Online

Direct links between dynamical, thermodynamic and structural properties of liquids: modelling results

L. Wang1, C. Yang1, M. T. Dove1, Yu. D. Fomin2, V. V. Brazhkin2, and K. Trachenko1 1 School of Physics and Astronomy, Queen Mary University of London, Mile End Road, London, E1 4NS, UK and 2 Institute for High Pressure Physics, RAS, 142190, Moscow, Russia

We develop an approach to liquid thermodynamics based on collective modes. We perform exten- sive molecular dynamics simulations of noble, molecular and metallic liquids and provide the direct evidence that liquid energy and specific heat are well-described by the dependence of the Frenkel (hopping) frequency. The agreement between predicted and calculated thermody- namic properties is seen in the notably wide range of temperature spanning tens of thousands of Kelvin. The range includes both subcritical liquids and supercritical fluids. We discuss the struc- tural crossover and inter-relationships between structure, dynamics and thermodynamics of liquids and supercritical fluids.

INTRODUCTION because the interactions are both strong and system- dependent, the resulting energy and other thermody- It is an interesting fact that the liquid state has proven namic functions will also be strongly system-dependent, to be difficult to describe by theory throughout the his- precluding their calculation in general form and under- tory of condensed matter research [1–12]. The problem standing using basic principles, in contrast to solids and extends beyond condensed matter and exists in other ar- gases [12]. Consistent with this somewhat pessimistic eas where strong interactions are combined with dynam- view, the discussion of liquid thermodynamic properties ical disorder such as field theory. has remained scarce. Indeed, physics textbooks have very little, if anything, to say about liquid specific heat, in- In a weakly-interacting system such as a dense gas, the cluding textbooks dedicated to liquids [1–12]. potential energy is much smaller than the kinetic energy. As recently reviewed [17], emerging evidence advances These systems are amenable to perturbation treatment our understanding of the thermodynamics of the liquid giving corrections to the non-interacting case [5]. Pertur- state. The start point is the early theoretical idea of J bation approaches have been widely explored to calculate Frenkel [1] who proposed that liquids can be considered liquid thermodynamic properties but have not been able as solids at times smaller than liquid relaxation time, τ, to agree with experiments. For example, the analysis of the average time between two particle rearrangements at tractable models such as van der Waals or hard-spheres one point in space. This implies that in liquids systems returns the gas-like result for the liquid constant- 3 will be similar to those in solids for frequencies above the volume specific heat cv = 2 kB [12–14]. This is in contrast Frenkel frequency ωF: to experimental results showing that cv of monatomic liq- uids close to the is nearly identical to the solid-like result, cv = 3kB and decreases to about 2kB 1 at high temperature [15, 16]. As expected on general ω > ωF = (1) τ grounds, the perturbation approach does not work for strongly-interacting systems. The above argument predicts that liquids are capa- Strong interactions are successfully treated in solids, ble of supporting shear modes, the property hitherto at- crystals or glasses, where the harmonic model is a good tributable to solids only, but only for frequencies above starting point and gives the most of the vibrational en- ωF. ergy. However, this approach requires fixed reference We note that low-frequency modes in liquids, sound points around which the energy expansion can be made. waves, are well-understood in the hydrodynamic regime

arXiv:1702.07221v1 [cond-mat.soft] 23 Feb 2017 With small vibrations around mean atomic positions, ωτ < 1 [18], however Eq. (1) denotes a distinct, solid- solids meet this requirement but liquids seemingly do not: like elastic regime of wave propagation where ωτ > 1. In liquid ability to flow implies that the reference lattice is essence, this suggests the existence of a cutoff frequency non-existent. ωF above which particles in the liquid can be described Therefore, liquids seemingly have no simplifying fea- by the same equations of motion as in, for example, solid tures such as small interactions of gases or small displace- glass. Therefore, liquid collective modes include both ments of solids [12]. In other words, liquids have no small longitudinal and transverse modes with frequency above parameter. One might adopt a general approach not re- ωF in the solid-like elastic regime and one longitudinal hy- lying on approximations and seek to directly calculate drodynamic mode with frequency below ωF (shear mode the liquid energy for a model system where interactions is non-propagating below frequency ωF as discussed be- and structure are known. This meets another challenge: low). 2

Recall the earlier textbook assertion [12] that a gen- collective modes in liquids play the same role, in view of eral thermodynamic theory of liquids can not be devel- the earlier Frenkel proposal and recent experimental evi- oped because liquids have no small parameter. How is dence? We have started exploring this question [37] just this fundamental problem addressed here? According to before the high-frequency transverse modes were directly Frenkel’s idea, liquids behave like solids with small oscil- measured and subsequently developed it in a number of lating particle displacements serving as a small param- ways [17]. This involves calculating the liquid energy eter. Large-amplitude diffusive particle jumps continue as the energy where transverse modes propagate to play an important role, but do not destroy the exis- above ωF in Eq. (1). tence of the small parameter. Instead, the jumps serve to The main aim of this paper is to provide direct com- modify the phonon spectrum: their frequency, ωF, sets putational evidence to the phonon theory of liquid ther- the minimal frequency above which the small-parameter modynamics and its predictions. We achieve this by cal- description applies and solid-like modes propagate. culating the liquid energy and ωF in extensive molecular It has taken a long time to verify this picture exper- dynamics simulations. In the next chapter, we briefly imentally. The experimental evidence supporting the discuss the main steps involved in calculating the liq- propagation of high-frequency modes in liquids currently uid energy. We then proceed to calculating the liquid includes inelastic X-ray, neutron and Brillouin scattering energy and Frenkel frequency independently from molec- experiments but most important evidence is recent and ular dynamics simulations using several methods which follows the deployment of powerful synchrotron sources agree with each other. We do this for three systems cho- of X-rays [19–36]. sen from different classes of liquids: noble, metallic and Early experiments detected the presence of high- molecular, and find good agreement between predicted frequency longitudinal acoustic propagating modes and and calculated results in the wide range of temperature mapped dispersion curves which were in striking resem- and pressure. The range includes both subcritical liq- blance to those in solids [19]. These and similar results uids and supercritical state below the Frenkel line where were generated at temperature just above the melting. transverse waves propagate. We calculate and analyze The measurements were later extended to high temper- liquid energy and cv using several different methods. Fi- atures considerably above the melting point, confirming nally, we discuss how our results offer insights into inter- the same result. It is now well established that liquids relationships between structure, dynamics and thermo- sustain propagating modes with wavelengths extending dynamics in liquids and supercritical fluids. down towards interatomic separations, comparable to the wave vectors of phonons in crystals at the Brillouin zone boundaries [20–32]. More recently, the same result has PHONON APPROACH TO LIQUID been asserted for supercritical fluids [24, 31, 32]. Impor- THERMODYNAMICS tantly, the propagating modes in liquids include acoustic transverse modes. These were first seen in highly viscous Calculating liquid energy fluids (see, e.g., Refs. [33, 34]), but were then studied in low-viscosity liquids on the basis of positive disper- We summarize the main result of calculation of the liq- sion [20–22, 25] (the presence of high-frequency trans- uid energy on the basis of propagating modes. A detailed verse modes increases sound velocity from the hydrody- discussion can be found in a recent review [17]. namic to the solid-like value). These studies included wa- According to the previous discussion, the propagating ter [35], where it was found that the onset of transverse modes in liquids include two transverse modes propagat- excitations coincides with the inverse of liquid relaxation ing in the solid-like elastic regime with frequency ω > ωF. time [36], as predicted by Frenkel [1]. The energy of these modes, together with the energy of More recently, high-frequency transverse modes in liq- the longitudinal mode gives the liquid vibrational energy. uids were directly measured in the form of distinct dis- In addition to vibrations, particles in the liquids undergo persion branches and verified on the basis of computer diffusive jumps between quasi-equilibrium positions as modeling [26–30], and the striking similarity between discussed above. Adding the energy of these jumps to dispersion curves in liquids and their crystalline (poly- the phonon energy in the Debye model gives the total crystalline) counterparts was noted. We note that the energy of thermal motion in the liquid [17, 37]: contribution of high-frequency modes is particularly im- portant for liquid thermodynamics because these modes  3! make the largest contribution to the energy due to ωF ET = NT 3 − (2) quadratic density of states. ωD The above discussion calls for an important ques- tion about liquid thermodynamics. In solids, collective where N is the number of particles and ωD is transverse modes, phonons, play a central role in the theory, in- Debye frequency and the subscript refers to thermal mo- cluding the theory of thermodynamic properties. Can tion. Here and below, kB = 1. 3

At low temperature, τ  τD, where τD is the Debye vibration period, or ωF  ωD. In this case, Eq. (2) 1 dE gaslike gives the specific heat cv = N dT close to 3, the solid-like result. At high temperature when τ → τD and ωF → ωD, Frenkel line e Eq. (2) gives c close to 2. The decrease of c from 3 to 2 r

v v u t a

with temperature is consistent with experimental results r liquidlike e in monatomic liquids [15, 16]. The decrease of cv is also p m seen in complex liquids [38]. e T

Eq. (2) attributes the experimental decrease of cv with temperature to the reduction of the number of transverse 1 Pressure modes above the frequency ωF = τ . The comparison of this effect with experiments can be more detailed if cv is compared in the entire temperature range where it de- creases from 3 to 2. This meets the challenge that ωF in Eq. (2) is not directly available in the cases of interest. ωF (τ) is measured is dielectric relaxation or NMR experi- FIG. 1: Colour online. The Frenkel line in the supercriti- ments in systems responding to electric or magnetic fields cal region. Particle dynamics includes both oscillatory and only. These liquids are often complex and do not include diffusive components below the line, and is purely diffusive simple model systems that are widely studied theoreti- above the line. Below the line, the system is able to sup- port rigidity and transverse modes at high frequency. Above cally such as liquid Ar. Importantly, the range of mea- the line, particle motion is purely diffusive, and the ability to sured ωF does not extend to high frequency comparable support rigidity and transverse modes is lost at all available to ωD, and it is in this range where liquid cv undergoes an frequencies. Crossing the Frenkel line from below corresponds important change from 3 to 2 as discussed above. ωF can to the transition between the “rigid” liquid to the “non-rigid” G∞ gas-like fluid. be calculated from the Maxwell relationship ωF = η , where G∞ is the instantaneous shear modulus and η is viscosity taken from a different experiment [17]. More recently, it has been suggested [40] that taking the shear excludes the state of the liquid where cv 3 modulus at a finite high frequency (rather than infinite can gradually reduce to 2 and where interesting physics frequency) agrees better with the modelling data. Apart operates. However, this becomes possible above the crit- from rare estimations [39, 40], G∞ is not available. In ical point. This brings us to the interesting discussion practice, the comparison of experimental cv and cv pre- of the supercritical state of matter. Theoretically, little dE dicted as dT with E given by Eq. (2) is done by keeping is known about the supercritical state, apart from the G∞ as a free parameter, obtaining a good agreement be- general assertion that supercritical fluids can be thought tween experimental and predicted cv and observing that of as high-density gases or high-temperature fluids whose G∞ lies in the range of several GPa typical for liquids properties change smoothly with temperature or pressure [17, 37]. In the last few years, Eq. (2) and its extensions and without qualitative changes of properties. This as- to include the phonon anharmonicity and quantum ef- sertion followed from the known absence of a phase tran- fects of phonon excitations was shown to account for the sition above the critical point. We have recently proposed experimental cv of over 20 different systems, including that this picture should be modified, and that a new line, metallic, noble, molecular and network liquids [17]. the Frenkel line (FL), exists above the critical point and In view of the persisting problem of liquid thermody- separates two states with distinct properties (see Figure namics, it is important to test Eq. (2) directly by linking 1) [13, 14, 41, 42]. Physically, the FL is not related to the liquid energy (cv) on one hand and ωF on the other the critical point and exists in systems where the critical and testing the theory in a precise way. This, together point is absent. with achieving consistency with other approaches to cal- The main idea of the FL lies in considering how the culate the liquid energy, is one of the objectives of this particle dynamics change in response to pressure and study. Importantly, this programme includes supercriti- temperature. Recall that particle dynamics in the liquid cal fluids as well as subcritical liquids, as discussed below. can be separated into solid-like oscillatory and gas-like diffusive components. This separation applies equally to supercritical fluids as it does to subcritical liquids. In- Thermodynamics of supercritical fluids deed, increasing temperature reduces τ, and each particle spends less time oscillating and more time jumping; in- If the system is below the critical point (see Figure creasing pressure reverses this and results in the increase 1), the temperature increase eventually results in boiling of time spent oscillating relative to jumping. Increasing and the first-order transition, with cv discontinuously de- temperature at constant pressure or density (or decreas- 3 creasing to about 2 in the gas phase. The intervening ing pressure at constant temperature) eventually results 4 in the disappearance of the solid-like oscillatory motion of partial charges derived using the Distributed Multipole particles; all that remains is the diffusive gas-like motion. Analysis method [46]. Fe was simulated using the many- This disappearance represents the qualitative change in body embedded-atom potential [47]. In the case of CO2, particle dynamics and gives the point on the FL in Fig- the electrostatic interactions were evaluated using the ure 1. Most important system properties qualitatively smooth particle mesh Ewald method. The MD systems change either on the line or in its vicinity [13, 14, 41, 42]. were first equilibrated in the constant pressure and tem- In a given system, the FL exists at arbitrarily high pres- perature ensemble at respective state points for 20 ps. sure and temperature, as does the melting line. System properties were subsequently simulated at differ- Quantitatively, the FL can be rigorously defined by ent and averaged in the constant energy pressure and temperature at which the minimum of the and volume ensemble for 30 ps. velocity autocorrelation function (VAF) disappears [14]. We are interested in properties of real dense strongly- Above the line defined in such a way, velocities of a large interacting liquids with potential energy comparable to number of particles stop changing their sign and particles kinetic energy and hence have chosen fairly high densi- lose the oscillatory component of motion. Above the line, ties: ρ = 1.5 g/cm3 and ρ = 1.9 g/cm3 for Ar, ρ = 8 VAF is monotonically decaying as in a gas [14]. For the g/cm3 and ρ = 11 g/cm3 for Fe and ρ = 1.34 g/cm3 purposes of this discussion, the significance of the FL is for CO2. The lowest temperature in each simulation was that the phonon approach to liquids and Eq. (2) apply the melting temperature at the corresponding density, to supercritical fluids below the FL to the same extent Tm. The highest temperature significantly exceeded the as they apply to subcritical liquids. Indeed, the presence temperature at the Frenkel line at the corresponding den- of an oscillatory component of particle motion below the sity, TF, taken from the earlier calculation of the Frenkel FL implies that τ is a well-defined parameter and that line in Ar [14], Fe [48] and CO2 [49]. As discussed above, transverse modes propagate according to Eq. (1). The the temperature range between Tm and TF corresponds ability of the supercritical system to sustain solid-like to the regime where transverse modes progressively dis- rigidity at frequency above ωF suggested the term “rigid” appear and where Eq. (2) applies. We have simulated liquid to differentiate it from the “non-rigid” gas-like fluid 100−700 temperature points at each pressure depending above the FL [13, 14]. on the system. The number of temperature points was Therefore, the FL separates the supercritical state into chosen to keep the temperature step close to 10 K. two states where transverse modes can and cannot prop- agate. This is supported by direct calculation of the cur- As discussed above, Eq. (2) applies to subcritical liq- rent correlation functions [43] showing that propagating uids as well as to supercritical fluids below the Frenkel and non-propagating transverse modes are separated by line. Accordingly, our simulations include the tempera- the Frenkel line. Interestingly, Eq. (2) can serve as a ture range both below and above the critical tempera- thermodynamic definition of the FL: the loss of the oscil- ture. This will be discussed in more detail below. latory component of particle motion at the FL approxi- mately corresponds to τ → τD (here, τD refers to Debye period of transverse modes) or ωF → ωD. According to Eq. (2), this gives cv of about 2. Using the criterion RESULTS AND DISCUSSION cv = 2 gives the line that is in remarkably good coin- cidence with the line obtained from the VAF criterion above [14]. Liquid energy and

We have calculated ωF in (2) from its definition in (1), SIMULATION DETAILS 1 as ωF = τ . τ can be calculated in a number of ways. Most common methods calculate τ as decay time of the We have considered liquids from three important sys- self-intermediate scattering or other functions by the fac- tem types: noble Ar, molecular CO2 and metallic Fe. tor of e or as the time at which the mean-squared dis- We have used the molecular dynamics (MD) simula- placement crosses over from ballistic to diffusive regime tion package DL POLY [44] and simulated systems with [50]. These methods give τ in agreement with a method 4576 − 8000 particles with periodic boundary conditions. employing the overlap function depending on the cutoff The interatomic potential for Ar is the pair Lennard- parameter ac provided ac = 0.3a, where a is the inter- Jones potential [45], known to perform well at elevated molecuar distance [50]. We use the latter method and pressure and temperature. For CO2 and Fe, we have calculate τ at 13-20 temperature points at each den- used interatomic potentials optimized tested in the liq- sity depending on the system. At each density, we fit uid state at high pressure and temperature. The po- τ to the commonly used Vogel-Fulcher-Tammann depen- 1 tential for CO2 is the rigid-body nonpolarizable poten- dence and use ωF = τ to calculate the liquid energy tial based on a quantum chemistry calculation, with the predicted from the theory. The predicted cv is calculated 5

1 dE as cv = N dT where E is given by Eq. (2):

 3 2 (a) ωF 3T ωF dωF 3000 Theory cv = 3 − − 3 (3) ωD ωD dT Simulation

TC where N is the number of atoms for Ar and Fe and the 2000 number of molecules for CO2. The first two terms in (3) give c = 2 when ω tends to 3.0 Theory v F Simulation Energy (K) 2.7 its high-temperature limit of ωF. The last term reduces N Fluctuation B

dωF 1000 /k cv below 2 by a small amount because is close to zero v 2.4 dT c at high temperature [17]. 2.1 T (K) We now compare the calculated energy and cv with 1.8 0 500 1000 1500 those directly computed in the MD simulations. We 0 note that the energy in Eq. (2) is the energy of ther- 200 400 600 800 1000 1200 1400 Temperature (K) mal phonon motion, ET , which contributes to the total liquid energy as

10000

E = E0 + ET (4) (b) Theory 8000 Simulation where E0 is liquid energy at zero temperature and repre- sents temperature-independent background contribution 6000 due to the interaction energy. In comparing the calculated ET in Eq. (2) with the 3.0 Theory 4000 Simulation energy from MD simulations, we therefore subtract the Energy (K) 2.7 N Fluctuation B

constant term from the MD energy. The comparison of /k

v 2.4 1 dE c cv = N dT is performed directly because the constant 2000 2.1 term does not contribute to cv. We have also calcu- T (K) 1.8 0 1500 3000 4500 lated cv using the fluctuations formula for the kinetic en- 2 2 0 ergy K in the constant energy ensemble: hK i − hKi = 500 1000 1500 2000 2500 3000 3500 4000 2 1.5(kBT ) N(1−1.5kB/cv) [51]. Both methods agree well, Temperature (K) as follows from Figures 2a and 2b. There is only one adjustable parameter in Eq. (2), ωD, which is expected to be close to transverse Debye frequency. ω is independently calculated from the MD FIG. 2: Energy per particle and specific heat of Ar at den- F sity ρ = 1.5 g/cm3 (a) and ρ = 1.9 g/cm3 (b). Solid and simulation as discussed above. In Figures 2 and 3 we dashed lines correspond to results from simulations and the- compare the energy and cv calculated on the basis of ory, respectively. The large (blue) circle corresponds to crit- Eqs. (2) and (3) and compare them with those computed ical temperature. The black solid curves in the insets show 1 dE in MD simulations. Blue circle in each figure shows the cv calculated as cv = N dT . Solid diamonds correspond to cv critical temperature. We observe good agreement be- calculated from the fluctuation formula (see text). The red tween predicted and calculated properties in a temper- (dashed) line is the theoretical result for cv. ature range including both subcritical and supercritical temperature. This involved using τD ≈ 0.6 ps (ρ = 8 3 3 g/cm ) and τD ≈ 0.2 ps (ρ = 11 g/cm ) for Fe, τD ≈ 0.9 The agreement between the predicted and calculated 3 3 ps (ρ = 1.5 g/cm ) and τD ≈ 0.3 ps (ρ = 1.9 g/cm ) results supports the interpretation of the decrease of cv for Ar and τD ≈ 0.5 ps for CO2, in reasonable order-of- with temperature discussed in the Introduction: ωF de- magnitude agreement with experimental τD of respective creases with temperature, and this causes the reduction crystalline systems as well as maximal frequencies seen of the number of transverse modes propagating above ωF in experimental liquid dispersion curves (see, e.g., [27]). and hence the reduction of cv. We note the expected trend of τD reducing with density. For CO2, the same mechanism operates except we need At high temperature where ωF ≈ ωD, Eq. (3) predicts to account for degrees of freedom in a molecular system. cv close to 2, noting that the last term gives only a small We first consider the case of solid CO2. The MD in- contribution to cv because ωF becomes slowly varying at teratomic potential treats CO2 molecules as rigid linear high temperature. Consistent with this prediction, we units, contributing the kinetic term of 2.5 to the specific observe the decrease of cv from 3 to 2 in Figures 2 and 3. heat per molecule including 1 from the rotational degrees 6

50000 3300 (a) Theory 3000 Theory 40000 Simulation Simulation T C 2500 TC 30000

3.3 5.2 Theory 2000 20000 3.0 Theory

Energy (K) Simulation Energy (K) 4.8 Simulation N

2.7 B N B /k 4.4

v /k

2.4 c v 1500 10000 c 2.1 4.0 T (K) T (K) 1.8 3.6 0 7500 15000 200 400 600 0 1000 5000 10000 15000 20000 200 300 400 500 600 700 800 Temperature (K) Temperature (K)

FIG. 4: Energy per particle and specific heat of CO2 at den- (b) sity ρ = 1.34 g/cm3. Solid and dashed lines correspond to 125000 Theory results from simulations and theory, respectively. The large Simulation (blue) circle corresponds to critical temperature. The black 100000 1 dE TC solid curves in the insets show cv calculated as cv = N dT . The red (dashed) line is the theoretical result for cv. 75000 3.3 Theory 3.0 Simulation Energy (K) 50000 N 2.7

B Consistent with the above discussion, we observe that /k v 2.4 c cv for CO2 calculated directly from the MD simulations 25000 2.1 is close to 5 at low temperature just above melting. At T (K) 1.8 0 20000 40000 60000 this temperature, ωF  ωD, giving the solid-like value of 0 12500 25000 37500 50000 cv as in the case of monatomic Ar and Fe. As temper- Temperature (K) ature increases, two transverse modes of inter-molecular motion progressively disappear, resulting in the decrease of cv to the value of about cv = 4, in agreement with cv calculated from the theoretical equation for ET. FIG. 3: Energy per particle and specific heat of Fe at density 3 3 We note that the temperature range in which we com- ρ = 8 g/cm (a) and ρ = 11 g/cm (b). Solid and dashed lines pare the predicted and calculated properties is notably correspond to results from simulations and theory, respec- tively. The large (blue) circle corresponds to critical temper- large (e.g., 200 − 8000 K for Ar, and 2000 − 55000 K for ature. The black solid curves in the insets show cv calculated Fe). This range is 10-100 times larger than those typi- 1 dE cally considered earlier [17]. The higher temperatures for as cv = N dT . The red (dashed) line is the theoretical result for cv. Fe might appear as unusual, however we note that liquid as well as supercritical iron fluid remains an unmod- ified system up to very high temperature: the first ion- of freedom of the linear molecular and 1.5 from transla- ization potential of Fe is 7.9 eV, or over 90,000 K. Hence the considered temperature range is below the tempera- tions (here, we have noted that CO2 molecules librate and rotate in the solid at low and high temperature, respec- ture at which the system changes its structure and type tively [52]). Noting the potential energy contributes the of interactions. same term due to equipartition, the specific heat becomes The very wide temperature range reported here is 5 per molecule. This implies that for molecular CO2, Eqs. mostly related to the large part of the temperature inter-   3 ωF val in Figures 2-4 being above the critical point where no (2) modifies as ET = NT 5 − , where N is the ωD phase transition intervenes and where the liquid phase number of molecules and ωF is related to the jump fre- exists at high temperature, in contrast to subcritical liq- quency of molecules and which gives cv = 5 in the solid uids where the upper temperature is limited by the boil- state where ωF is infinite. We use the modified equation ing line. The agreement between predicted and calcu- to calculate the energy and cv and compare them to those lated properties in such a wide temperature range adds computed from the MD simulation in Figure 4. support to the phonon approach to liquid thermodynam- 7 ics we propose. lution of collective modes is qualitatively different below We make three points regarding the observed agree- and above the FL [17]. Below the line, transverse modes ment between the calculated and predicted results. First, disappear starting from the lowest frequency ωF. Above the collective modes contributing to the thermal energy the line, the remaining longitudinal mode starts disap- 2πc in (2) are considered to be harmonic. The anharmonic- pearing starting from the highest frequency L , where ity can be accounted for in the Gr¨uneisenapproxima- L is the particle mean free path (no oscillations can take tion, however this involves an additional parameter [17]. place at distance smaller than L). This gives qualita- We attempted to avoid introducing additional parame- tively different behavior of the energy and cv below and ters and sought to test Eq. (2) which contains only one above the FL, resulting in their crossover at the FL [17]. parameter, ωD. Interestingly, the thermodynamic crossover at cv = 2 Second, Eq. (2) involves the Debye model and implies a structural crossover. Indeed, the energy per quadratic density of states (DOS). This approximation particle in a system with pair-wise interactions is is justified since the Debye model is particularly rele- vant for disordered isotropic systems such as glasses [12], ∞ which are known to be nearly identical to liquids from the 3 Z E = k T + 4πρ r2U(r)g(r)dr (5) structural point of view. Furthermore, the experimental 2 B dispersion curves in liquids are very similar to those in 0 solids such as poly-crystals [28–30]. Therefore, the De- bye model can be used in liquids to the same extent as where ρ = N/V is number density and g(r) is radial in solids. One important consequence of this is that the distribution function. high-frequency range of the phonon spectrum makes the According to Eq. (4), the liquid energy is E = E0+ET, largest contribution to the energy, as it does in solids in- where ET is given by Eq. (2). If the system energy un- cluding disordered solids. We also note that liquid DOS dergoes the crossover at the FL where cv = 2, Eq. (5) can be represented as the sum of solid-like and gas-like implies that g(r) should also undergo a crossover. There- components in the two-phase thermodynamic model [53], fore, the structural crossover in liquids can be predicted and the solid-like component can be extracted from the on the basis of the thermodynamic properties. liquid DOS calculated in MD simulations. This can pro- We also expect the structural crossover at the FL vide more information about the DOS beyond Debye ap- to be related to the dynamical crossover on general proximation. grounds. As discussed above, below the FL particles os- Third, Eq. (2) assumes a lower frequency cutoff for cillate around quasi-equilibrium positions and occasion- 1 ally jump between them. The average time between transverse waves, ωF = τ , as envisaged by Frenkel in (1). Our recent detailed analysis of the Frenkel equa- jumps is given by liquid relaxation time, τ (Figure 5 tions shows that the dispersion relationship for liquid schematically shows a local jump event from its surround- q ing “cage”.) This means that a static structure exists transverse modes is ω = c2k2 − 1 , where c is the s 4τ 2 s during τ for a large number of particles below the FL, shear speed of sound and k is wavenumber [17]. Here, giving rise to the well-defined medium-range order com- ω gradually crosses over from 0 to its solid-like branch parable to that existing in disordered solids [55]. On the 1 ω = csk when ω  ωF = τ . In this sense, using a lower other hand, the particles lose the oscillatory component frequency cutoff in (2) might be thought of as an approx- of motion above the FL and start to move in a purely imation. However, we have recently shown [54] that the diffusive manner as in gases. This implies that the fea- square-root dependence of ω gives the liquid energy that tures of g(r) are expected to be gas-like. As a result, is identical to (2). g(r) medium-range peaks are expected to have different temperature dependence below and above the FL. This behavior was observed in Ar in MD simulations in the Structural crossover and its relationship to short-range structure [56]. More recently, the crossover dynamical and thermodynamic properties in supercritical Ne in the medium range at the FL was ascertained on the basis of X-ray scattering experiments The results in the previous sections support the picture [57]. in which the decrease of liquid cv from 3 to 2 is related to In Figure 6a we plot pair distribution functions (PDFs) reduction of the energy of transverse modes propagating of Ar at density ρ = 1.9 g/cm3 in a wide temperature above ωF as described by Eq. (3). According to Eq. (3), range. Using the FL criterion cv = 2 gives the temper- cv = 2 corresponds to complete disappearance of trans- ature at the FL, TF, of about 4000 K at that density, verse modes at the FL when ωF ≈ ωD (the disappearance which we find to be consistent with the criterion of the is supported by the direct calculation of transverse modes disappearance of the minimum of the velocity autocorre- on the basis of current correlation functions [43]). Impor- lation function [14]. The PDF was calculated with the tantly, cv = 2 marks the crossover of cv because the evo- distance step of 0.05 A,˚ giving 600 PDF points. 8

3.5 (a) 3.0 250 K 500 K 2.5 1000 K 2.0 2000 K FIG. 5: Schematic representation of a jump event in the liq- 4000 K uid. g(r) 1.5 8000 K

1.0 We observe PDF peaks in the medium range order up to about 20 A˚ at low temperature. The peaks reduce and 0.5 broaden with temperature. To study this in more detail, 0.0 we plot the peak heights vs temperature in Figure 6b. We 5 10 15 20 25 30 observe that the medium-range third and fourth peaks persist well above the critical temperature (Tc = 151 K r (Å) for Ar): the highest temperature simulated corresponds to 53Tc. This interestingly differs from the traditional ex- pectation that the structure of the matter so deep in the 100 supercritical state has gas-like features only. At temper- (b) ature above TF, the height of the fourth peak becomes h-1 comparable to its temperature fluctuations (calculated 10-1 as the standard deviation of the peak height over many structures separated in time by 1 ps at each tempera- ture) by order of magnitude. The fifth and higher-order 10-2 peaks disappear before the highest temperature in the simulated range is reached. nd We plot the peak heights in Figure 6b in the double- 10-3 2 peak logarithmic plot because we expect to see an approximate 3rd peak power-law decay of the peak heights at low temperature. th Indeed, PDF in solids can be represented as a set of Gaus- 4 peak T (K) 10-4 sian functions with peaks heights h depending on tem- 103 104 perature as h ∝ √1 exp − α  where α is a temperature- T T independent factor [1, 58]. This temperature dependence of h was also quantified in MD simulations [59]. h de- crease mostly due to the factor √1 whereas the effect of T FIG. 6: (a) Pair distribution functions of Ar at different tem- the exponential factor on h is small and serves to reduce peratures. The temperatures correspond to the first peak the rate at which h decrease [59]. This implies that in decreasing from top to bottom at 250 K, 500 K, 1000 K, 2000 solids, log h ∝ − log T approximately holds. K, 4000 K and 8000 K; (b) h − 1, where h are the heights of In liquids, we expect the same relationship to hold be- PDF peaks at different temperatures. The lines are linear fits to the low-temperature data range. low the FL where τ  τD, corresponding to a particle oscillating many times before diffusively moving to the next quasi-equilibrium position. Indeed, the ratio of the number of diffusing particles Ndif to the total number of generally small compared to the harmonic energy. This Ndif τD particles N in the equilibrium state is N = τ [17] at is witnessed by the closeness of high-temperature cv to Ndif its harmonic result for both solids and high-temperature any given moment of time. N is small when τ  τD be- low the FL and can be neglected. Hence, log h ∝ − log T liquids [60, 61]. applies to liquids at any given moment of time below the We therefore expect that log(h−1) ∝ − log T approxi- FL where τ  τD. This also applies to longer observa- mately holds in the low-temperature range below the FL tion times if h is averaged over τ [17]. We note that the as in solids but deviates from the linearity around the above result, h ∝ √1 , involves the assumption that the T crossover at the FL where τ → τD and where the dy- energy of particle displacements is harmonic (see, e.g., namics becomes gas-like (the calculated PDF in Fig. 6a Ref. [1]). Anharmonicity becomes appreciable at high is normalized to 1 where no correlations are present at temperature, however the anharmonic energy terms are large distances; hence we plot h − 1 in order to compare 9 it with the theoretical result h ∝ √1 which tends to [4] N. H. March and M. P. Tosi, Atomic Dynamics in Liquids T zero when no correlations are present at high tempera- (Dover Publications, 1991). ture). We note that the crossover is expected to be broad [5] S. Chapman and T. G. Cowling, The Mathematical The- ory of Non-uniform gases (Cambridge University Press, because τ  τD applies well below the FL only. A sub- 1995). stantial diffusive motion takes place in the vicinity of the [6] U. Balucani and M. Zoppi, Dynamics of the Liquid State Ndif line where N can not be neglected, affecting the linear (Oxford University Press, 1995). relationship. [7] J. M. Ziman, Models of disorder, (Cambridge University Consistent with the above prediction, we observe the Press, 1995). linear regime at low temperature in Figure 6b, followed [8] R. Zwanzig, Non-Equlibrium Statistical Mechanics (Ox- ford University Press, 2001). by the deviation from the straight lines taking place [9] J. L. Barrat and J. P. Hansen, Basic Concepts for Simple around 3000 K for the 2nd peak, 5000 K for the 3rd and Complex Fluids (Cambridge University Press, 2003). peak and 4000 K for the 4th peak, respectively. The [10] T. E. Faber, Introduction to the Theory of Liquid Metals smooth crossover in the 3000-5000 K range is centered (Cambridge University Press, 2010). around 4000 K, consistent with the temperature at the [11] J. P. Hansen and I. R. McDonald, Theory of Simple Liq- Frenkel line discussed above. We also note that 4000 K uids (Elsevier, 2013). corresponds to the specific heat c = 2 in Figure 2b, in [12] L. D. Landau and E. M. Lifshitz, Statistical Physics v (Pergamon Press, 1969). agreement with the earlier discussion. [13] V. V. Brazhkin, Yu. D. Fomin, A. G. Lyapin, V. N. Ryzhov and K. Trachenko, Phys. Rev. E 85, 031203 (2012). [14] V. V. Brazhkin, Yu. D. Fomin, A. G. Lyapin, V. N. SUMMARY Ryzhov, E. N. Tsiok and K. Trachenko, Phys. Rev. Lett. 111, 145901 (2013). As discussed in the Introduction, liquids have been [15] G. Grimvall, Phys. Scripta 11, 381 (1975). viewed as inherently complicated systems lacking useful [16] D. C. Wallace, Phys. Rev. E 57, 1717 (1998). theoretical concepts such as a small parameter [12]. To- [17] K. Trachenko and V. V. Brazhkin, Rep. Prog. Phys. 79, 016502 (2016). gether with recent experimental evidence and theory [17], [18] L. D. Landau and E. M. Lifshitz, Fluid Mechanics the modelling data presented here and its quantitative (Butterworth-Heinemann, 1987). agreement with predictions are beginning to change this [19] J. R. D. Copley and J. M. Rowe, Phys. Rev. Lett. 32, 49 traditional perspective. Our extensive molecular dynam- (1974). ics simulations of liquid energy and specific heat provide [20] W. C. Pilgrim, S. Hosokawa, H. Saggau, H. Sinn and E. direct evidence for the link between dynamical and ther- Burkel, J. Non-Cryst. Sol. 250-252, 96 (1999). modynamic properties of liquids. We have found this to [21] E. Burkel, Rep. Prog. Phys. 63, 171 (2000). [22] W. C. Pilgrim and C. Morkel J. Phys.: Condens. Matter. be the case for several important types of liquids at both 18, R585 (2006). subcritical and supercritical conditions spanning thou- [23] T. Scopigno, G. Ruocco and F. Sette, Rev. Mod. Phys. sands of Kelvin. This supports an emerging picture that 77, 881 (2005). liquid thermodynamics can be understood on the basis [24] G. Ruocco and F. Sette, J. Phys.: Condens. Matt. 11, of high-frequency collective modes. A more general im- R259 (1999). plication is that, contrary to the prevailing view, liquids [25] W. C. Pilgrim and C. Morkel, J. Phys.: Cond. Matt. 18, are emerging as systems amenable to theoretical under- R585 (2006). [26] S. Hosokawa et al Phys. Rev. Lett. 102, 105502 (2009). standing in a consistent picture as is the case in solid [27] S. Hosokawa, M. Inui, Y. Kajihara, S. Tsutsui and A. Q. state theory. In addition to the link between dynamical R. Baron, J. Phys.: Condens. Matt. 27, 194104 (2015). and thermodynamic properties, we have discussed how [28] V. M. Giordano and G. Monaco, PNAS 107, 21985 these properties are related to liquid structure. (2010). This research utilised MidPlus computational facilities [29] V. M. Giordano and G. Monaco, Phys. Rev. B 84, 052201 supported by QMUL Research-IT and funded by the EP- (2011). [30] S. Hosokawa et al, J. Phys.: Condens. Matt. 25, 112101 SRC Grant EP/K000128/1. We acknowledge the support (2013). of the Royal Society, RFBR (15-52-10003) and CSC. [31] A. Cunsolo et al, J. Chem. Phys. 114, 2259 (2001). [32] G. G. Simeoni et al, Nat. Phys. 6, 503 (2010). [33] M. Grimsditch, R. Bhadra and L. M. Torell, Phys. Rev. Lett. 62, 2616 (1989). [34] F. Scarponi, L. Comez, D. Fioretto and L. Palmieri, [1] J. Frenkel, Kinetic Theory of Liquids (Oxford University Phys. Rev. B 70, 054203 (2004). Press, 1947). [35] E. Pontecorvo et al, Phys. Rev. E 71, 011501 (2005). [2] J. P. Boon and S. Yip, Molecular Hydrodynamics [36] A. Cunsolo et al Phys. Rev. B 85, 174305 (2012). (Courier Dover Publications, 1980). [37] K. Trachenko, Phys. Rev. B 78, 104201 (2008). [3] N. H. March, Liquid Metals (Cambridge University [38] A. R. Dexter and A. J. Matheson, Trans. Faraday Soc. Press, 1990). 64, 2632 (1968). 10

[39] D. C. Wallace, Phys. Rev. A 25, 3290 (1982). J. Chem. Phys. 135, 194503 (2011). [40] F. Puosi and D. Leporini, J. Chem. Phys. 136, 041104 [51] D. Frenkel and B. Smit, Understanding molecular simu- (2012). lation: from algorithms to applications (Academic press, [41] V. V. Brazhkin and K. Trachenko, Physics Today 2001). 65(11), 68 (2012). [52] S. Liu, M. A. Doverspike and M. S. Conradi, J. Chem. [42] V. V. Brazhkin, A. G. Lyapin, V. N. Ryzhov, K. Tra- Phys. 81, 6064 (1984). chenko, Yu. D. Fomin, E. N. Tsiok, Physics-Uspekhi [53] S. Lin, M. Blanco and W. A. Goddard III, J. Chem. Phys. 55(11), 1061 (2012). 119, 11792 (2003). [43] Yu. D. Fomin, V. N. Ryzhov, E. N. Tsiok, V. V. Brazhkin [54] C. Yang et al (unpublished). and K. Trachenko, J. Phys.: Condens. Matt 28, 43LT01 [55] P. S. Salmon, R. A. Martin, P. E. Mason and G. J. Cuello, (2016). Nature 435, 75 (2005). [44] I. T. Todorov, B. Smith, M. T. Dove and K. Trachenko, [56] D. Bolmatov, V. V. Brazhkin, Yu. D. Fomin, V. N. J. Mater. Chem. 16, 1911 (2006). Ryzhov and K. Trachenko, J. Chem. Phys. 139, 234501 [45] J. L. Yarnell, M. J. Katz, R. G. Wenzel and S. H. Koenig, (2013). Phys. Rev. A 7, 2130 (1973). [57] C. Prescher et al, arXiv:1608.08437. [46] M. Gao, A. M. Misquitta, C. Yang, I. T. Todorov, [58] A. A. Maradudin, E. W. Montroll, G. H. Weiss and I. A. Mutters and M. T. Dove (unpublished); M. Gao, P. Ipatova, Theory of Lattice Dynamics in the Harmonic PhD thesis “ absorption in metal organic Approximation (Academic, New York, 1971). frameworks”. [59] S. O. Yurchenko, N. P. Kryuchkov and A. V. Ivlev, J. [47] A. B. Belonoshko, R. Ahuja and B. Johansson, Phys. Chem. Phys. 143, 034506 (2015). Rev. Lett. 84, 3638 (2000). [60] E. I. Andritsos et al, J. Phys. Condens. Matt. 25, 235401 [48] Y. Fomin et al, Sci. Rep. 4, 7194 (2014). (2013 ). [49] C. Yang, V. V. Brazhkin, M. T. Dove and K. Trachenko, [61] D. Bolmatov and K. Trachenko, Phys. Rev. B 84, 054106 Phys. Rev. E 91, 012112 (2015). (2011). [50] S. Sengupta, F. Vasconcelos, F. Affouard and S. Sastry,