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

Effect of characteristic size on the collective phonon transport in crystalline GeTe

Kanka Ghosh [email protected] University of Bordeaux, CNRS, Arts et Metiers Institute of Technology, Bordeaux INP, INRAE, I2M Bordeaux, 351 Cours de la libération, F-33400 Talence, France    Andrzej Kusiak University of Bordeaux, CNRS, Arts et Metiers Institute of Technology, Bordeaux INP, INRAE, I2M Bordeaux, 351 Cours de la libération, F-33400 Talence, France    Jean-Luc Battaglia University of Bordeaux, CNRS, Arts et Metiers Institute of Technology, Bordeaux INP, INRAE, I2M Bordeaux, 351 Cours de la libération, F-33400 Talence, France
Abstract

We study the effect of characteristic size variation on the phonon thermal transport in crystalline GeTe for a wide range of temperatures using the first-principles density functional method coupled with the kinetic collective model approach. The characteristic size dependence of phonon thermal transport reveals an intriguing collective phonon transport regime, located in between the ballistic and the diffusive transport regimes. Therefore, systematic investigations have been carried out to describe the signatures of phonon hydrodynamics via the competitive effects between grain size and temperature. A characteristic non-local length, associated with phonon hydrodynamics and a heat wave propagation length has been extracted. The connections between phonon hydrodynamics and these length scales are discussed in terms of the Knudsen number. Further, the scaling relation of thermal conductivity as a function of characteristic size in the intermediate size range emerges as a crucial indicator of the strength of the hydrodynamic behavior. A ratio concerning normal and resistive scattering rates has been employed to understand these different scaling relations, which seems to control the strength and prominent visibility of the collective phonon transport in GeTe. This systematic investigation emphasizes the importance of the competitive effects between temperature and characteristic size on phonon hydrodynamics in GeTe, which can lead to a better understanding of the generic behavior and consequences of the phonon hydrodynamics and its controlling parameters in low-thermal conductivity materials.

I Introduction

Detecting phonon hydrodynamics and associated collective phonon transport in low-thermal conductivity materials is a challenging task due to the not-so-overwhelming differences between the normal and the resistive phonon scattering rates. This leads to the exploration of very low cryogenic temperatures to see a visible effect of collective motion of phonons. On the other hand, 2D materials draw an appreciable amount of studies [1, 2, 3] concerning phonon hydrodynamics because of their enhanced normal scattering phenomena. This helps in realizing phonon hydrodynamics even at higher temperatures and therefore can be understood using experiments. Nevertheless, collective phonon transport holds fundamental interest in materials as it draws parallel to the hydrodynamic flow in fluids. Investigating this collective phonon transport in low-thermal conductivity materials is crucial to understand the role of different competing effects that influences phonon hydrodynamics and invokes fundamental question on its generality and validity in both high and low conductivity materials. The complete understanding of the origin of this phenomena thus demands a systematic decoupling between various controlling parameters that dictate phonon hydrodynamics in materials.

Phonon hydrodynamics is a heat transport phenomena where the collective flow of phonons dominate the heat conduction in materials [4, 5, 1, 6, 7, 8]. This is enabled by significantly higher momentum conserving normal scattering (N) events compared to other dissipative scattering events [Umklapp (U), isotope (I) and boundary scattering (B)], favoring damped wave propagation of temperature fluctuations [9, 10]. In their consecutive two pioneering theoretical works [11, 12] published in 1966, Guyer and Krumhansl distinguished the phonon hydrodynamics for nonmetallic crystals using the comparison between normal and resistive average scattering rates. Phonon hydrodynamics have also been realized by the deviation from Fourier’s law at certain length and time scales [13, 2]. The concept of the kinetic theory of relaxons to characterize phonon hydrodynamics have been introduced by Cepellotti and Marzari [14]. Very recently, Sendra et al. [15] introduced a framework to use hydrodynamic heat equations from phonon Boltzmann equation to study the hydrodynamic effects in semiconductors.

Experiments and theoretical investigations over the years suggest that only few and mostly two-dimensional (2D) materials possess phonon hydrodynamics [5, 2, 3, 16, 17]. Some of these 2D materials like graphene and boron nitride [1] can even persist phonon hydrodynamics at room temperature due to the presence of strong normal scattering realized via first-principles simulations. Recently, the relation between the thickness and thermal conductivity and consequently their connection to the phonon hydrodynamics was studied for graphite [18]. The presence of second sound, a prominent manifestation of phonon hydrodynamics, was also observed in graphite at a temperature higher than 100 K via the experiments carried out by Huberman et al. [19]. This validates the predictions of the simulation studies done by Ding et al. [20] on graphite. Similarly, theoretical evaluations by Markov et al. [9] confirmed the experimental observation [21] of hydrodynamic Poiseuille phonon flow in bismuth (Bi) at low temperature. A faster than T3T^{3} scaling of the lattice thermal conductivity was described as a marker to identify phonon hydrodynamics in bulk black phosphorus [8] and SrTiO3 [22, 23]. Koreeda et al. [24] studied collective phonon transport in KTaO3 using low frequency light-scattering and time-domain light-scattering techniques and phonon hydrodynamics was found to exist below 30 K . Further second sound was also observed in solid helium (0.6 - 1 K) [25], NaF (\sim 15 K) [26] at low temperatures.

As discussed earlier, the studies of phonon hydrodynamics for low-thermal conductivity materials are substantially less compared to its high-thermal conductivity counterpart. However, a systematic decoupling of various controlling parameters can help manipulate phonon hydrodynamic behavior in the low-thermal conductivity materials. Torres et al. [27] showed a strong phonon hydrodynamic behavior in low-lattice thermal conductivity (κL\kappa_{L}) materials such as single layer transition metal dichalcogenides (MoS2, MoSe2, WS2 and WSe2). In our earlier paper [28], we investigated the low temperature thermal transport in crystalline GeTe, a chalcogenide-based material of diverse practical interests [29, 30], which shows even lower lattice thermal conductivity compared to metal dichalcogenides and found that it exhibits phonon hydrodynamics. However we found that the presence of hydrodynamic phonon transport in crystalline GeTe is sensitive to the grain size and vacancies present in the material. Further, temperature was found to play an important role in favoring appreciable normal scattering events to enable collective phonon transport.

For low-thermal conductivity materials like GeTe, the characteristic size of the material and temperature are two crucial parameters that influence the existence of phonon hydrodynamics. Distinguishing the competing effects of these two factors is important for general understanding of collective phonon transport in GeTe. Therefore, in the current paper, we investigate the effects of characteristic size (LL) on the collective thermal transport in low-thermal conductivity crystalline GeTe for temperatures ranging from 4 K to around 500 K. We use first-principles calculations with a kinetic collective model approach [31] for this paper. We first identify the LL-regimes corresponding to ballistic and complete diffusive regimes. Then we explore the regime of collective phonon transport that comprises both ballistic and diffusive phonons. Average scattering rates have been used to identify phonon hydrodynamic regimes both in terms of temperature and characteristic size. Further, temperature and LL-regimes are quantified using the Knudsen number obtained using two different length scales concerning phonon hydrodynamics. The prominent signature of phonon hydrodynamics in GeTe is found to depend on the scaling exponent of thermal conductivity as a function of LL in the intermediate LL-regime where phonon transport shifts from ballistic to complete diffusive. The ratio of normal to resistive scattering rates at this LL-regime seems to dictate the strength of the hydrodynamic behavior.

II Computational Details

First-principles density functional methods are employed to optimize the structural parameters of crystalline GeTe (space group R3mR3m). The details of the parameters for GeTe can be found in our earlier paper [32]. The phonon lifetime is calculated using PHONO3PY [33] software package. The supercell approach with finite displacement of 0.03 Å is employed to obtain the harmonic (second order) and the anharmonic (third order) force constants, given via

Φαβ(lκ,lκ)=2Φuα(lκ)uβ(lκ)\Phi_{\alpha\beta}(l\kappa,l^{\prime}\kappa^{\prime})=\frac{\partial^{2}\Phi}{\partial u_{\alpha}(l\kappa)\partial u_{\beta}(l^{\prime}\kappa^{\prime})} (1)

and

Φαβγ(lκ,lκ,l′′κ′′)=3Φuα(lκ)uβ(lκ)uγ(l′′κ′′)\Phi_{\alpha\beta\gamma}(l\kappa,l^{\prime}\kappa^{\prime},l^{\prime\prime}\kappa^{\prime\prime})=\frac{\partial^{3}\Phi}{\partial u_{\alpha}(l\kappa)\partial u_{\beta}(l^{\prime}\kappa^{\prime})\partial u_{\gamma}(l^{\prime\prime}\kappa^{\prime\prime})} (2)

respectively. Density functional method is implemented with QUANTUM-ESPRESSO [34] to calculate the forces acting on atoms in supercells. The harmonic force constants are approximated as [33]

Φαβ(lκ,lκ)Fβ[lκ;u(lκ)]uα(lκ)\Phi_{\alpha\beta}(l\kappa,l^{\prime}\kappa^{\prime})\simeq-\frac{F_{\beta}[l^{\prime}\kappa^{\prime};\textbf{u}(l\kappa)]}{u_{\alpha}(l\kappa)} (3)

where F[ll^{\prime}κ\kappa^{\prime}; u(llκ\kappa)] is atomic force computed at r(ll^{\prime} κ\kappa^{\prime}) with an atomic displacement u(lκl\kappa) in a supercell. Similarly, third order force constants are calculated using[33]

Φαβγ(lκ,lκ,l′′κ′′)Fγ[l′′κ′′;u(lκ),u(lκ)]uα(lκ)uβ(lκ)\Phi_{\alpha\beta\gamma}(l\kappa,l^{\prime}\kappa^{\prime},l^{\prime\prime}\kappa^{\prime\prime})\simeq-\frac{F_{\gamma}[l^{\prime\prime}\kappa^{\prime\prime};\textbf{u}(l\kappa),\textbf{u}(l^{\prime}\kappa^{\prime})]}{u_{\alpha}(l\kappa)u_{\beta}(l^{\prime}\kappa^{\prime})} (4)

where F[l′′l^{\prime\prime}κ′′\kappa^{\prime\prime}; u(llκ\kappa), u(ll^{\prime} κ\kappa^{\prime})] is the atomic force computed at r(l′′l^{\prime\prime} κ′′\kappa^{\prime\prime}) with a pair of atomic displacements u(lκl\kappa) and u(lκl^{\prime}\kappa^{\prime}) in a supercell. These two sets of linear equations are solved using the Moore-Penrose pseudoinverse as is implemented in PHONO3PY [33].

Using the 2×\times2×\times2 supercell and finite displacement method, we obtain 228 supercell configurations with different pairs of displaced atoms, for the calculations of the anharmonic force constants. A larger 3×\times3×\times3 supercell is employed for the harmonic force constants calculation. For force calculations, the reciprocal space is sampled with a 3×\times3×\times3 k-sampling Monkhorst-Pack (MP) mesh [35] shifted by a half-grid distances along all three directions from the Γ\Gamma- point. For the density functional calculations, the Perdew-Burke-Ernzerhof (PBE) [36] generalized gradient approximation (GGA) is used as the exchange-correlation functional. Due to its negligible effects on the vibrational features of GeTe, as mentioned in earlier studies [37, 38], the spin-orbit interaction has been ignored. Electron-ion interactions are represented by pseudopotentials using the framework of the projector-augmented-wave (PAW) method [39]. The Kohn-Sham (KS) orbitals are expanded in a plane-wave (PW) basis with a kinetic cutoff of 60 Ry and a charge density cutoff of 240 Ry as specified by the pseudopotentials of Ge and Te. The total energy convergence threshold has been kept at 10-10 a.u. for supercell calculations. The imaginary part of the self-energy has been calculated using the tetrahedron method from which phonon lifetimes are obtained.

III Lattice dynamics and Kinetic Collective Model (KCM)

In the theory of lattice dynamics, the crystal potential is expanded with respect to atomic displacements and the third-order coefficients associated with anharmonicity are used to calculate the imaginary part of the self-energy [33]. Generally, in a harmonic approximation, phonon lifetimes are infinite whereas, anharmonicity in a crystal yields a phonon self-energy Δωλ\Delta\omega_{\lambda} + iΓλi\Gamma_{\lambda}. The phonon lifetime (τphph\tau_{ph-ph}) has been computed from the imaginary part of the phonon self energy as τλ\tau_{\lambda} = 12Γλ(ωλ)\frac{1}{2\Gamma_{\lambda}(\omega_{\lambda})} using PHONO3PY [33, 40] from the following equation

Γλ(ωλ)=18π2λλ′′Δ(q+q+q′′)Φλλλ′′2{(nλ+nλ′′+1)δ(ωωλωλ′′)+(nλnλ′′)[δ(ω+ωλωλ′′)δ(ωωλ+ωλ′′)]}\Gamma_{\lambda}(\omega_{\lambda})=\frac{18\pi}{\hbar^{2}}\sum_{\lambda^{\prime}\lambda^{\prime\prime}}\Delta\left(\textbf{q}+\textbf{q}^{\prime}+\textbf{q}^{\prime\prime}\right)\mid\Phi_{-\lambda\lambda^{\prime}\lambda^{\prime\prime}}\mid^{2}\{(n_{\lambda^{\prime}}+n_{\lambda^{\prime\prime}}+1)\delta(\omega-\omega_{\lambda^{\prime}}-\omega_{\lambda^{\prime\prime}})+(n_{\lambda^{\prime}}-n_{\lambda^{\prime\prime}})[\delta(\omega+\omega_{\lambda^{\prime}}-\omega_{\lambda^{\prime\prime}})-\delta(\omega-\omega_{\lambda^{\prime}}+\omega_{\lambda^{\prime\prime}})]\} (5)

where nλn_{\lambda} = 1exp(ωλ/kBT)1\frac{1}{exp(\hbar\omega_{\lambda}/k_{B}T)-1} is the phonon occupation number at the equilibrium. Δ(q+q+q′′)\Delta\left(\textbf{q}+\textbf{q}^{\prime}+\textbf{q}^{\prime\prime}\right) = 1 if q+q+q′′=G\textbf{q}+\textbf{q}^{\prime}+\textbf{q}^{\prime\prime}=\textbf{G}, or 0 otherwise. Here G represents reciprocal lattice vector. Integration over q-point triplets for the calculation is made separately for normal (G = 0) and umklapp processes (G \neq 0) and therefore phonon umklapp (τU\tau_{U}) and phonon normal lifetime (τN\tau_{N}) have been distinguished. Using second-order perturbation theory, the scattering of phonon modes by randomly distributed isotopes (τI1\tau_{I}^{-1}) is given by Tamura [41] as

1τλI(ω) = πωλ22N∑_λδ(ω- ω’_λ ) ∑_k g_k—∑_αW_α(k,λ)W_α^*(k,λ)— ^2

(6)

where gkg_{k} is the mass variance parameter, defined as

gk=ifi(1mikm¯k)2g_{k}=\sum_{i}f_{i}\left(1-\frac{m_{ik}}{\overline{m}_{k}}\right)^{2} (7)

fif_{i} is the mole fraction, mikm_{ik} is the relative atomic mass of iith isotope, m¯k\overline{m}_{k} is the average mass = ifimik\sum_{i}f_{i}m_{ik}, and W is a polarization vector. The database of the natural abundance data for elements [42] is used for the mass variance parameters. The phonon-boundary scattering has been implemented using Casimir diffuse boundary scattering [43] as τλB\tau_{\lambda}^{B} = Lvλ\frac{L}{\mid\textbf{v}_{\lambda}\mid}, where, vλ\textbf{v}_{\lambda} is the average phonon group velocity of phonon mode λ\lambda and LL is the grain size, which is also called Casimir length, the length phonons travel before the boundary absorption or re-emission [43].

We use the kinetic collective model (KCM) [31] to obtain the lattice thermal conductivity of GeTe. The KCM method has emerged as a useful approach to depict heat transport at all length scales with the computational cost being substantially less than that of the full solution of the linearized Boltzmann transport equation. According to the KCM method, the heat transfer process occurs via both collective phonon modes, emerges from the normal scattering events and via independent phonon collisions. Therefore, lattice thermal conductivity can be expressed as a sum of both kinetic and collective contributions weighed by a switching factor (Σ[0,1]\Sigma\in\left[0,1\right]), which indicates the relative weight of normal and resistive scattering processes [31, 27]. While each mode exhibits individual phonon relaxation time in the kinetic contribution, the collective contribution is designated by an identical relaxation time for all modes [44, 31]. In the kinetic contribution term, the boundary scattering is included via the Matthiessen’s rule as

τk1=τU1+τI1+τB1\tau_{k}^{-1}=\tau_{U}^{-1}+\tau_{I}^{-1}+\tau_{B}^{-1} (8)

where τk\tau_{k} is the total kinetic phonon relaxation time. On the contrary, a form factor FF, calculated from the sample geometry, is employed to incorporate boundary scattering in the collective term [31, 44]. The KCM equations are:

κL=κk+κc\kappa_{L}=\kappa_{k}+\kappa_{c} (9)
κk=(1Σ)ωfTv2τkDdω\kappa_{k}=(1-\Sigma)\int\hbar\omega\frac{\partial f}{\partial T}v^{2}\tau_{k}D\textit{d}\omega (10)
κc=(ΣF)ωfTv2τcDdω\kappa_{c}=(\Sigma F)\int\hbar\omega\frac{\partial f}{\partial T}v^{2}\tau_{c}D\textit{d}\omega (11)
Σ=11+τNτRB\Sigma=\frac{1}{1+\frac{\langle\tau_{N}\rangle}{\langle\tau_{RB}\rangle}} (12)

where κk\kappa_{k} and κc\kappa_{c} are kinetic and collective contributions to κL\kappa_{L}, respectively. τN\langle\tau_{N}\rangle and τRB\langle\tau_{RB}\rangle designate average normal phonon lifetime and average resistive (considering UU, II, and BB) phonon lifetimes, respectively. τN\langle\tau_{N}\rangle and τRB\langle\tau_{RB}\rangle are defined in the KCM [31] as integrated mean-free times,

τRB=C1τk𝑑ωC1𝑑ω\langle\tau_{RB}\rangle=\frac{\int C_{1}\tau_{k}d\omega}{\int C_{1}d\omega} (13)

and

τN=C0τN𝑑ωC0𝑑ω\langle\tau_{N}\rangle=\frac{\int C_{0}\tau_{N}d\omega}{\int C_{0}d\omega} (14)

where τk\tau_{k} is the total kinetic relaxation time and phonon distribution function in the momentum space, represented in terms of Ci=0,1(ω)C_{i=0,1}(\omega), defined in Ref. [31] as

Ci(ω)=(v|q|ω)2iωfTDC_{i}(\omega)=\left(\frac{v|q|}{\omega}\right)^{2i}\hbar\omega\frac{\partial f}{\partial T}D (15)

where v(ω)v(\omega) is the phonon mode velocity and q\mid q\mid is modulus wave vector. C0C_{0} represents the specific heat of mode ω\omega. ff stands for Bose-Einstein distribution function, vv is mode velocity and D(ω)D(\omega) is phonon density of states for each mode. τc\tau_{c} denotes the total collective phonon relaxation time and defined as

τc(T)=C1𝑑ω(τI1+τU1)C1𝑑ω\tau_{c}(T)=\frac{\int C_{1}d\omega}{\int(\tau_{I}^{-1}+\tau_{U}^{-1})C_{1}d\omega} (16)

Σ\Sigma stands for the switching factor. FF is the form factor approximated via [44]

F(Leff)=Leff22π2l2(1+4π2l2Leff21)F(L_{eff})=\frac{L_{eff}^{2}}{2\pi^{2}l^{2}}\left(\sqrt{1+\frac{4\pi^{2}l^{2}}{L_{eff}^{2}}}-1\right) (17)

where, LeffL_{eff} is the effective length of the sample (in our system, we use LeffL_{eff} = LL, the grain size) and ll is the characteristic non-local scale [11, 44]. This characteristic non-local length ll emerges from the complete hydrodynamic description of the KCM and is defined as a parameter that determines the non-local range in phonon transport. In our earlier paper [28], comparing the results for thermal conductivity obtained using both direct solutions of linearized Boltzmann transport equation (LBTE) and KCM for GeTe, we found an excellent agreement between them at low temperature. At higher temperatures, a reasonable matching trend is retrieved, with KCM exhibiting slightly lower values than the LBTE solutions. However, in the low temperature hydrodynamic regime for GeTe, the solutions of LBTE and KCM collapse satisfactorily. For all KCM [31] calculations of lattice thermal conductivity and associated parameters, KCM.PY code [31] is implemented with the outputs obtained using PHONO3PY [33].

IV Results and Discussions

IV.1 Ballistic and diffusive phonon transport

As a first step to elucidating the complex collective behavior of phonons as a function of characteristic size (LL), it is imperative to explore the variation of κL\kappa_{L} with LL and therefore to identify the effect of LL on the ballistic and diffusive phonon transport. Figure 1 describes this

Refer to caption
Figure 1: The variation of lattice thermal conductivity (κL\kappa_{L}) as a function of characteristic length (LL) of the GeTe sample at different temperatures.

variation of GeTe for a wide temperature range (4 - 500 K). As the LL varies almost 10610^{6} orders of magnitude (from 0.001 μ\mum to 5000 μ\mum), κL\kappa_{L} undergoes a transition from a linear variation of LL to a plateau-like regime, and corresponds to complete ballistic and complete diffusive transport respectively. As we gradually go from lower to higher temperatures, the ballistic regime shrinks and the diffusive regime starts growing. Also, the onset of diffusive transport gradually seems to take place at lower

Refer to caption
Figure 2: The variation of lattice thermal conductivity (κL\kappa_{L}) with temperature as a function of LL for crystalline GeTe.
Refer to caption
Figure 3: The variation of (a) LballL_{ball} and (b) LdiffL_{diff} are represented as a function of temperature for crystalline GeTe. The insets of (a) and (b) display the defining procedure of LballL_{ball} and LdiffL_{diff} respectively for a representative case of TT = 10 K.

values of LL as we increase the temperature. It is well known in the literature [13, 45] that ballistic conduction of phonons occurs without ph-ph scattering and displays a linear variation with LL, whereas diffusive conduction of phonons manifests when scattered phonons carry the heat. The effect of the characteristic size on κL\kappa_{L} can also be represented via the variation of κL\kappa_{L} with temperatures for different LL, as shown in Fig 2. At higher temperatures, it is well known [32] that κL\kappa_{L} decreases with TT, with 1/TT scaling due to the dominant umklapp scattering at high temperatures. As temperature is lowered, gradually κL\kappa_{L} attains a peak following a gradual decrement at very low temperature. As we go towards higher LL, the peaks of κL\kappa_{L} as a function of temperature are gradually seen to be shifted towards lower temperatures (Fig. 2).

The effect of LL on the temperature variation of κL\kappa_{L} gives rise to an interesting feature as we increase LL above a certain limit. It is known that LL plays a crucial role via phonon-boundary scattering as gradual increment of LL assists in weakening the boundary scattering. This weakening of boundary scattering and strong normal scattering rates (to be discussed later) at low temperatures transforms the peak of κL\kappa_{L} into a cusp-like feature when LL \geq 1 μ\mum and κL\kappa_{L} is further seen to be increased at very low temperatures.

To give a more precise account of ballistic and diffusive conduction of phonons in GeTe, we further investigate the characteristic size range of ballistic and diffusive conduction as a function of temperature. The complete ballistic length regime (LballL_{ball}) is defined via the maximum value of LL, until which κL\kappa_{L} varies linearly with LL. Similarly, the complete diffusive length regime (LdiffL_{diff}) is defined via the minimum length LL, above which κL\kappa_{L} reaches the thermodynamic limit and therefore reaches a plateau. In other words, LdiffL_{diff} represents the longest mean free path of the heat carriers at a particular temperature [13]. Figures 3.(a) and 3.(b) represent the variations of LballL_{ball} and LdiffL_{diff} respectively, as a function of temperature. As temperature increases, we see a gradual decrement of both LballL_{ball} and LdiffL_{diff}. We note here that at very high temperatures, we hardly observe any ballistic conduction of phonons and the LdiffL_{diff} acquires a very low value. This is representative of the fact that at high temperatures, internal phonon-phonon scattering is so dominant that no ballistic heat conduction is seen to exist, even for very small grains of the order of 1 nm.

To delve deeper into the origin of length dependent κL\kappa_{L} in the ballistic phonon conduction regime of GeTe, we investigate the contribution of acoustic and optical modes in the ballistic propagation of heat. Earlier, molecular dynamics simulations and experiments on suspended single-layer graphene [46, 47] suggested the ballistic propagation of long-wavelength, low-frequency acoustic phonon to be solely responsible for the length-dependent κL\kappa_{L} in the ballistic regime. Our previous studies on GeTe [32, 28] suggested that GeTe shows a clear distinction between acoustic and optical modes in the frequency domain around 2.87 THz. The density of states goes to zero around a frequency of 2.87 THz [32], distinguishing two distinct frequency regimes: acoustic regime (ω\omega << 2.87 THz) and optical regime (ω\omega >> 2.87 THz). We calculate the cumulative lattice thermal

Refer to caption
Figure 4: The variation of cumulative lattice thermal conductivity (κLc\kappa_{L}^{c}) of crystalline GeTe as a function of phonon frequency (ω\omega) for four different temperatures: (a) 10 K, (b) 30 K, (c) 50 K and (d) 300 K. For each temperature, the variation of κLC\kappa_{L}^{C} with ω\omega is presented for three different LL: 0.001, 0.002 and 0.003 μ\mum. The gray shaded region denotes the acoustic modes regime for GeTe.

conductivity (κLc\kappa_{L}^{c}) as a function of phonon frequency defined as [33, 40]

κLc=0ωκL(ω)𝑑ω\kappa_{L}^{c}=\int_{0}^{\omega}\kappa_{L}(\omega^{\prime})d\omega^{\prime} (18)

where κL\kappa_{L} (ω\omega^{\prime}) is defined as [33, 40]

κL(ω)1NV0λCλvλvλτλδ(ωωλ)\kappa_{L}(\omega^{\prime})\equiv\frac{1}{NV_{0}}\sum_{\lambda}C_{\lambda}\textbf{v}_{\lambda}\otimes\textbf{v}_{\lambda}\tau_{\lambda}\delta(\omega^{\prime}-\omega_{\lambda}) (19)

with 1N\frac{1}{N} λδ(ωωλ)\sum_{\lambda}\delta(\omega^{\prime}-\omega_{\lambda}) the weighted density of states (DOS). Figure 4 presents the variation of average κLc\kappa_{L}^{c} (= (κxxC+κyyC+κzzC)3\frac{(\kappa_{xx}^{C}+\kappa_{yy}^{C}+\kappa_{zz}^{C})}{3}) with phonon frequency. The density of states goes to zero at a frequency where κLc\kappa_{L}^{c} reaches a plateau defining the separation between acoustic (frequency << 2.87 THz) and optical (frequency >> 2.87 THz) modes. Except at low temperature (TT = 10 K), the contribution from optical modes seem to present at all other temperatures. As we gradually increase the temperature [from Fig 4.(b) to Fig 4.(d)], the contributions from optical modes are seen to be enhanced. For example, for LL = 0.003 μ\mum, the contribution of optical modes at TT = 10 K, 30 K, 50 K and 300 K are 0 %\%, 9.9 %\%, 24.2 %\%, and 37.7 %\% respectively. Therefore, contrary to the understanding of ballistic propagation for a 2D material like single-layer graphene, except for very low temperatures, GeTe also shows a weak contribution from optical modes in the ballistic phonon propagation regime. However, the significant contributions come from acoustic modes in this regime.

To visualize the consequences on the mean-free path of the phonons at small LL, we present the variation of the effective mean-free path variation with phonon frequency for different LL at different temperatures in Fig. 5. In the KCM nomenclature, the kinetic mean free path [lk(ω)l_{k}(\omega)] and the collective mean free path [lc(T)l_{c}(T)] are defined as lk(ω)l_{k}(\omega) = vτkv\tau_{k} and lc(T)l_{c}(T) = v¯τc\overline{v}\tau_{c} respectively, where vv is the group velocity and

v¯=vωfTD(ω)𝑑ωωfTD(ω)𝑑ω\overline{v}=\frac{\int v\hbar\omega\frac{\partial f}{\partial T}D(\omega)d\omega}{\int\hbar\omega\frac{\partial f}{\partial T}D(\omega)d\omega} (20)

is the mean phonon velocity [31]. As the kinetic MFP

Refer to caption
Figure 5: Effective mean free path (MFP) of crystalline GeTe are presented as a function of frequencies for three different LL: 0.001 μ\mum (blue points), 0.002 μ\mum (green points) and 0.003 μ\mum (red points) at four different temperatures: (a) TT = 10 K, (b) TT = 30 K, (c) TT = 50 K, (d) and TT = 300 K. The gray shaded region denotes the acoustic modes regime for GeTe.

is a function of phonon frequency whereas the collective MFP is frequency independent and varies only with temperature, we present an effective MFP as leff(ω)l_{eff}(\omega) = (1-Σ\Sigma)lk(ω)l_{k}(\omega)+Σ\Sigma lcl_{c}. The separate contributions from collective and kinetic MFPs are described in Supplemental Fig. S1. Two effects can be observed from this representation. First, at low temperature, as the LL is increased, the optical modes at higher frequencies exhibit more scattered mean-free paths. Figure 5.(a) shows that at TT = 10 K, at higher frequencies in the optical modes, LL = 0.003 μ\mum persists more scattered MFPs compared to the LL = 0.001 μ\mum case. This feature indicates that the ballistic conduction is stronger for LL = 0.001 μ\mum, where LL strongly controls the mean-free path than that of the LL = 0.003 μ\mum case. Second, increasing temperature for fixed LL, also leads to the gradual weakening of the ballistic conduction of phonons, as can be seen from Fig. 5. This is evident from the gradual broadening of MFPs with temperature [follow fixed color points for four different temperatures in Fig. 5.(a), (b), (c) and (d).] due to the gradually weakening control of LL on dictating the mean-free paths of the system.

IV.2 Collective phonon transport

After understanding the effect of characteristic size (LL) on the ballistic and diffusive conduction of phonons, we turn our attention to the effect of LL on the collective phonon transport of crystalline GeTe. The connection between ballistic and diffusive phonon transport and the collective motion of phonons are crucial to determine the origin of the exotic hydrodynamic phonon transport in materials. In our earlier work [28], unusually, low-thermal conductivity chalcogenide GeTe emerged as a possible candidate to feature phonon hydrodynamics with the characteristic size being a dominant factor.

Refer to caption
Figure 6: Thermodynamic average phonon scattering rates as a function of temperature for GeTe for different characteristic sizes (LL). NN, UU, II and RR denote normal, umklapp, isotope and resistive scattering respectively. Boundary scattering rates for different LL are also presented. The shaded regions in (a) and (b) correspond to the validation of the Guyer’s condition [12] for Poiseuille’s flow (Eq. 24) for LL = 0.08 μ\mum and LL = 0.8 μ\mum respectively.
Refer to caption
Figure 7: The spectral representation of lattice thermal conductivity (κL\kappa_{L}) as a function of phonon frequency at TT = 10 K for four different characteristic size or grain sizes (LL): (a) 0.2 μ\mum, (b) 0.5 μ\mum, (c) 1 μ\mum and (d) 5 μ\mum. The kinetic contribution (κkinetic\kappa_{kinetic}) is defined using light violet and the collective contribution (κcollective\kappa_{collective}) is defined using light brown color.

In this context, we start by investigating the relative strengths of the average phonon scattering rates, which is defined as

τi1ave=λCλτλi1λCλ\langle\tau_{i}^{-1}\rangle_{ave}=\frac{\sum_{\lambda}C_{\lambda}\tau_{\lambda i}^{-1}}{\sum_{\lambda}C_{\lambda}} (21)

Here, λ\lambda denotes phonon modes (q, jj) comprising wave vector q and branch jj. Index ii denotes normal, umklapp, isotope, and boundary scattering processes used, marked by N, U, I, and B respectively. CλC_{\lambda} is the modal heat capacity, given by the following equation

Cλ=kB(ωλkBT)2exp(ωλ/kBT)[exp(ωλ/kBT)1]2C_{\lambda}=k_{B}\left(\frac{\hbar\omega_{\lambda}}{k_{B}T}\right)^{2}\frac{exp(\hbar\omega_{\lambda}/k_{B}T)}{[exp(\hbar\omega_{\lambda}/k_{B}T)-1]^{2}} (22)

where, TT denotes temperature, \hbar is the reduced Planck’s constant, and kBk_{B} is the Boltzmann constant. In one of the earliest works on phonon hydrodynamics, Guyer and Krumhansl [12] found that the hydrodynamic regime exists if

τU1aveτN1ave{}\langle\tau_{U}^{-1}\rangle_{ave}\ll\langle\tau_{N}^{-1}\rangle_{ave} (23)

Further, Guyer’s condition [12] for the presence of second sound and Poiseuille’s flow reads:

τU1ave<τB1ave<τN1ave{}\langle\tau_{U}^{-1}\rangle_{ave}<\langle\tau_{B}^{-1}\rangle_{ave}<\langle\tau_{N}^{-1}\rangle_{ave} (24)

In Fig 6, we explore the LL window that satisfies the aforementioned Guyer and Krumhansl condition of phonon hydrodynamics in crystalline GeTe. Figure 6 presents the average scattering rates due to normal (N), resistive (R) [comprised of umklapp (U) and isotope scattering (I)] and the phonon-boundary scattering as a function of temperature for GeTe. We observe a substantial width of LL, that persists phonon hydrodynamic conditions, as the boundary scattering rates decrease gradually on increasing LL. This is shown via the gray shaded regions in Figs. 6.(a) and 6.(b) for two representative grain sizes: LL = 0.08 μ\mum and LL = 0.8 μ\mum, respectively. In the scattering rate approach, we also identified the ballistic conduction region, mentioned earlier through the linear dependence of κL\kappa_{L} with LL, as the region where τB1aveτphph1ave\langle\tau_{B}^{-1}\rangle_{ave}\gg\langle\tau_{ph-ph}^{-1}\rangle_{ave}. Similarly, the purely diffusive conduction region, mentioned earlier as the LL-regime where κL\kappa_{L} is independent of LL, as the region where τB1aveτphph1ave\langle\tau_{B}^{-1}\rangle_{ave}\ll\langle\tau_{ph-ph}^{-1}\rangle_{ave}. At this point, we go back to Fig. 2 to explain the cusp-like behavior of κL\kappa_{L} as a function of temperature. This cusp-like pattern of κL\kappa_{L} is found to present for LL >> 1 μ\mum, as we gradually decrease the temperature. In Fig. 6.(b), this LL regime is identified as LL values for which normal scattering overpowers boundary scattering rates. At low temperatures, umklapp scattering is rare and boundary scattering acts as the dominant resistive phonon scattering. So, the effect of boundary scattering tries to reduce the κL\kappa_{L} while the momentum conserving normal scattering tries to increase κL\kappa_{L}. Overpowering normal scattering compared to boundary scattering for LL >> 1 μ\mum forces κL\kappa_{L} to increase after an apparent shallow dip or a plateau and gives rise to the cusp-like pattern in Fig. 2.

Once the Guyer and Krumhansl conditions are satisfied and a prominent LL window is observed to feature phonon hydrodynamics, we next investigate the spectral representation of lattice thermal conductivity (κL\kappa_{L}) in this LL window. In Fig. 7, using the KCM approach, we present a spectral representation of κL\kappa_{L}, distinguished by its kinetic (κkinetic\kappa_{kinetic}) and collective contributions (κcollective\kappa_{collective}), as a function of phonon frequency at TT = 10K for four different LL. We choose TT = 10 K as a representative temperature to feature collective transport of phonons. The four different LL values have been chosen such that it covers a wide range that traverses from ballistic transport to the hydrodynamic regime at TT = 10 K. As we gradually increase the LL [from Figs. 7(a) to 7(d)), a gradual increment of the contributions coming from the collective transport is observed (shown via the red shaded regions inside the curve). The spectral κL\kappa_{L} goes to zero before 2.87 THz, indicating the sole contribution of acoustic phonons in thermal transport at 10 K, as was realized earlier in Fig. 4(a).

To quantify the collective motion as a function of temperature for different LL, we investigate the variation of characteristic non-local length (ll) in GeTe at different temperatures and grain sizes. In a complete hydrodynamic description of thermal transport, the extension of the Guyer and Krumhansl equation [11] done in the KCM framework [44], namely, the hydrodynamic KCM equation, yields

τdQdt+Q=κT+l2(2Q+2Q)\tau\frac{\textit{d}\textbf{Q}}{\textit{dt}}+\textbf{Q}=-\kappa\nabla T+l^{2}\left(\nabla^{2}\textbf{Q}+2\nabla\nabla\cdot\textbf{Q}\right) (25)

where τ\tau is the total phonon relaxation time, Q is the heat flux, κ\kappa is phonon thermal conductivity, and ll is the non-local length, that determines the non-local range in

Refer to caption
Figure 8: The variation of Knudsen number (Kn) with temperature for different LL values of crystalline GeTe. The shaded region satisfies 0.1 \leq Kn \leq 10 while the rectangular boxes define phonon hydrodynamic regimes calculated from average scattering rates. Blue dashed lines to guide the eye for TT = 6 K, 10 K and 20 K.

phonon transport. If we employ the steady state, strong geometric effects, and neglect the term 2Q2\nabla\nabla\cdot\textbf{Q}, then the equation can be visualized as analogous to Navier-Stokes equation with ll resembling heat viscosity. The Knudsen number (Kn) can be obtained from Kn=l/LKn=l/L to study the collective motion quantitatively. Lower values of Kn define the recovery of Fourier’s law whereas the hydrodynamic behavior becomes prominent when Kn gets higher values [44, 7]. Figure 8 presents the variation of Kn as a function of temperature for different LL. As temperature is lowered, a gradual increment of Kn is observed, concomitant with the gradual prominence of non-local behavior. Kn has earlier been described [7, 9] to indicate a phonon hydrodynamic regime when 0.1 \leq Kn \leq 10, bearing similarities with fluid hydrodynamics. We denote this region via a shaded region in Fig. 8. In Fig. 8, we also superpose the hydrodynamic LL-window, identified using average scattering rates following Guyer and Krumhansl conditions for three representative temperatures: TT = 6 K, 10 K and 20 K. We observe that both definitions match well and the hydrodynamic LL-window obtained by scattering rate analysis falls within the Kn range for hydrodynamics.

Knudsen number calculation also reveals the Ziman hydrodynamic regime for GeTe. Looking at the vertical dashed lines corresponding to TT = 6 K and TT = 10 K in Fig. 8, a small LL-region is observed which does not fall into the rectangles, defined to indicate a hydrodynamic regime using scattering rate hierarchy. However, they fall inside the regime of 0.1 << Kn << 10, especially in the regime where Kn is close to 0.1. This corresponds to the Ziman hydrodynamic regime which corresponds to a regime where N scattering dominates but dissipates mostly by R scattering contrary to the Poiseuille hydrodynamic regime where N scattering dissipates mostly by the boundary scattering of the phonons. On the other hand, looking at LL values that lie inside 0.1 << Kn << 10 but with values close to 10, also sometimes do not lie inside the rectangular region (see the case of LL = 0.04 and 0.1 μ\mum at TT = 10 K in Fig 8). Recalling Fig 6.(a), we observe that LL = 0.04 μ\mum at TT = 10 K designates a scattering rate hierarchy, where τB1ave>τN1ave>τR1ave\langle\tau_{B}^{-1}\rangle_{ave}>\langle\tau_{N}^{-1}\rangle_{ave}>\langle\tau_{R}^{-1}\rangle_{ave}, but τB1ave\langle\tau_{B}^{-1}\rangle_{ave} is not \gg τphph1ave\langle\tau_{ph-ph}^{-1}\rangle_{ave}. Therefore, though it follows the prescribed hierarchy for hydrodynamics, the LL values do not enable a complete ballistic propagation and a competition between ballistic and diffusive phonons operates. This competition makes it difficult to distinguish sharp boundaries between different regimes. We will discuss more about this competition later. To characterize the repopulation of phonons in a different way, following Markov et al. [9], we extract a length scale related to the propagation of heat wave before being dissipated, called the heat wave propagation length (LhwplL_{hwpl}), defined as a length at which the completely diffusive thermal conductivity decays 1/e times:

κL(T,L)L=Lhwpl=κL(T,L>Ldiff)/e\kappa_{L}(T,L)\mid_{L=L_{hwpl}}=\kappa_{L}(T,L>L_{diff})/e (26)
Refer to caption
Figure 9: Heat wave propagation length (LhwplL_{hwpl}) as a function of temperature for crystalline GeTe. The temperature variation of phonon propagation lengths, correspond to the damping due to resistive scattering (λhydro\lambda_{hydro}) and both resistive and normal scattering (λgas\lambda_{gas}) along a and hexagonal c-axis are also presented.

where LdiffL_{diff} is the minimum length LL, above which κL\kappa_{L} reaches the thermodynamic limit, as mentioned earlier in Fig. 3. LhwplL_{hwpl} is connected to second sound, a typical characteristic for hydrodynamic heat transport phenomenon, which demonstrates the heat propagation as damped waves in a crystal [12, 1, 48] as a result of coherent collective motion of phonons due to the domination of N scattering. In this context, drift velocity of phonons (v¯\overline{v}) and phonon propagation length (λph\lambda_{ph}) are defined as

v¯j2=αCα𝐯αjg𝐯αjgαCα\overline{v}_{j}^{2}=\frac{\sum_{\alpha}C_{\alpha}\mathbf{v}_{\alpha j}^{g}\cdot\mathbf{v}_{\alpha j}^{g}}{\sum_{\alpha}C_{\alpha}} (27)

and

λph=v¯/τ1ave\lambda_{ph}=\overline{v}/\langle\tau^{-1}\rangle_{ave} (28)

where, CαC_{\alpha} is heat capacity of mode α\alpha, 𝐯αjg\mathbf{v}_{\alpha j}^{g} is phonon group velocity of mode α\alpha and jj can be either the component along the aa axis (xx) or the hexagonal cc axis (zz). Heat transfer of GeTe is anisotropic, as can be recalled from our earlier studies [32, 28], featuring different group velocities along the hexagonal cc axis and its perpendicular (aa axis) direction and therefore yields different drift velocities and different phonon propagation lengths along xx and zz. Figure 9 presents the variation of heat wave propagation length (LhwplL_{hwpl}) with temperature along with the variation of phonon propagation lengths along xx and zz. Phonon propagation lengths are distinguished [9] as λhydro\lambda_{hydro} and λgas\lambda_{gas} via

λhydro=v¯/τR1ave\lambda_{hydro}=\overline{v}/\langle\tau_{R}^{-1}\rangle_{ave} (29)
λgas=v¯/(τR1ave+τN1ave)\lambda_{gas}=\overline{v}/\left(\langle\tau_{R}^{-1}\rangle_{ave}+\langle\tau_{N}^{-1}\rangle_{ave}\right) (30)

Figure 9 shows the variation of heat wave propagation length (LhwplL_{hwpl}), superimposed with phonon propagation lengths with temperature along both aa and cc axis directions of GeTe. We observe that LhwplL_{hwpl} follows well the trend of λhydro\lambda_{hydro} as a function of temperature in the whole temperature range studied. λgas\lambda_{gas}, the phonon propagation length corresponds to the uncorrelated phonon gas where both N and R scattering processes contribute to the damping of heat wave, on the other hand, seems to diverge from LhwplL_{hwpl} as the temperature is lowered. This feature is an indication of gradual prominence of hydrodynamic behavior of phonons as the temperature is lowered. Similarly, the reasonable match between LhwplL_{hwpl} and λhydro\lambda_{hydro} predicts that heat wave propagation length is well captured by phonon flow with resistive damping caused by umklapp and isotope scattering. At very low temperature (TT = 4 K), a slight deviation is observed between LhwplL_{hwpl} and λhydro\lambda_{hydro} which can be attributed to the importance of boundary scattering as a significant damping process at very low temperature.

Therefore, LhwplL_{hwpl} can lead to the identification of the length scale at different temperatures at which phonon hydrodynamics can exist and therefore Poiseuille’s flow and second sound phenomena can be observed. Interestingly, comparing LhwplL_{hwpl} and characteristic size (LL) of the sample, we can define Knudsen number in another approach as [9] Kn = LhwplL_{hwpl}/LL. The variation of Kn obtained using LhwplL_{hwpl}, is presented as a function of temperature in the Supplemental Material (Fig. S2). The variation of Kn with TT is found similar to our earlier evaluation of Kn using nonlocal length (Fig. 8).

The blurry regions of transitions between ballistic, hydrodynamic, and diffusive transport are intriguing to understand the competition between different phonons with a wide range of mean free paths. Ideally, phonons with a wide spectrum of mean-free paths can be distinguished as either ballistic (MFP >> LL) or diffusive (MFP << LL) phonons. However, the relative strength between ballistic and diffusive phonons are crucial to realize the competition between these two kind of phonons which eventually plays a decisive role in dictating the visible hydrodynamic phenomena. The phonon Knudsen minimum is such an indicator for the transition between ballistic and hydrodynamic phonon propagation regimes and had been used for several materials including graphene [16], graphite [20], SrTiO3 [23], black phosphorus [18] to detect phonon hydrodynamics. Figures 10.(a) and (b) present the the variation of normalized thermal conductivity (κL\kappa_{L}^{*} = κL/L\kappa_{L}/L), a quantity that is similar to dimensionless κL\kappa_{L}, as a function of inverse Knudsen number, calculated using nonlocal length and heat wave propagation lengths respectively. Figure 10.(b) shows a wider range of 1/Kn as the Kn obtained using heat wave propagation length reaches higher values at low temperatures compared to that of the non-local length calculation from hydrodynamic KCM method. However, we observe almost similar trends of κL\kappa_{L}^{*} with the variation of 1/Kn coming out of the two different approaches in obtaining the Knudsen number. At TT = 300 K, a steep linear decreasing trend is observed which is associated with the diffusive phonon scattering events as phonons behave as uncorrelated gas particles and resistive scattering is prominent and dominating at this temperature.

Refer to caption
Figure 10: (a) The variation of normalized thermal conductivity (κL/L\kappa_{L}/L) as a function of inverse Knudsen number, calculated using characteristic non-local length for different temperatures. (b) The variation of normalized thermal conductivity (κL/L\kappa_{L}/L) as a function of inverse Knudsen number, calculated using heat wave propagation length for different temperatures.
Refer to caption
Figure 11: The variation of lattice thermal conductivity (κL\kappa_{L}) as a function of characteristic length (LL) in log-log scale for different temperatures: (a) TT = 4 K, (b) TT = 6 K and (c) TT = 10 K. The inset of Fig 11.(b) refers to the zoomed in view around linear to superlinear scaling at TT = 6 K. The intermediate regimes, located in between the ballistic and diffusive propagation regimes are shown via gray shades.

Starting from TT = 20 K, a gradual onset of a horizontal regime is visible before the linearly decreasing trend of κL\kappa_{L}^{*} as the temperature is lowered. At TT = 4 K, surprisingly, a cusp-like trend, resembling that of a shallow minimum followed by a prominent maximum is observed before a linearly decreasing κL\kappa_{L}^{*} at higher 1/Kn. The cusp-like shallow minimum at TT = 4 K indicates the phonon Knudsen minimum and predicts the presence of prominent transition from ballistic to hydrodynamic regime. Further, a prominent maximum in κL\kappa_{L}^{*} has only been observed at TT = 4 K, which designates the strong presence of hydrodynamic phonon transport in GeTe. Similar observation can be found by Li et al. [16] for suspended graphene, where the increasing trend of κL\kappa_{L}, normalized by sample width, was attributed to the strong presence of hydrodynamic phonon transport.

The behavior of phonon Knudsen minimum of GeTe convinces us to understand the competition between ballistic and diffusive phonons in the quasi-ballistic regimes of phonon transport. We specifically turn our attention toward the reason behind the strong presence of hydrodynamics at TT = 4 K visible through Knudsen minimum in Fig 10. We recall that even TT = 6 K, TT = 8 K persist in phonon hydrodynamics, realized via the average scattering rate analysis and Knudsen number variation with temperature. To perceive the reason behind the difference between strong and weak phonon hydrodynamics, we investigate the scaling relation between κL\kappa_{L} and LL in the intermediate regime of transport, where the transport is neither fully ballistic nor fully diffusive.

Figure 11 describes the variation of κL\kappa_{L} with LL at TT = 4 K, 6 K, and 10 K. Three phonon propagation regimes have been identified. At lower values of LL, ballistic phonons dominate the transport and therefore a linear dependency of κL\kappa_{L} on LL is observed. At very high LL, the phonon transport is completely diffusive and a plateau-like regime is observed, denoting an independence of κL\kappa_{L} over LL. The intermediate regime where the phonon propagation shifts from complete ballistic to complete diffusive, plays a crucial role in determining the strong or weak presence of hydrodynamic propagation of phonons. Figure 11.(c) indicates a sublinear variation in the intermediate regime at TT = 10 K. At TT = 6 K [Fig. 11(b)], a minute superlinear behavior is observed while at TT = 4 K [Fig. 11(a)], an enhanced superlinear behavior is perceived in the intermediate regime.

In the intermediate quasi-ballistic regime of phonon propagation, where both ballistic and diffusive phonons

Refer to caption
Figure 12: Variation of the scaling exponent α\alpha as a function of LL for different temperatures. The black dashed line denotes the α\alpha = 1 line.

operate and compete with each other, seems to be a marker to designate sample sizes (LL) with strong hydrodynamic phonon transport characteristics. To further quantify the intermediate nonliearity (both sub and superlinearity), we evaluate and present the scaling exponent[20] α\alpha = log(κL)\partial log(\kappa_{L})/log(L)\partial log(L) as a function of LL for different temperatures in Fig. 12.

α\alpha = 0 indicates the size-independent behavior of κL\kappa_{L} and therefore describes the completely diffusive phonon propagation regime. On the other hand, α\alpha = 1 reveals the linear size dependency and henceforth the complete ballistic phonon conduction regime. The superlinear dependence of κL\kappa_{L} on LL in the intermediate regime can be captured by the the condition α\alpha >> 1. From Fig. 12 we observe that at low LL, for low temperatures, α\alpha goes to 1. for higher temperatures, as expected almost no ballistic regime is observed with α\alpha << 1. As we increase LL, in the intermediate regime, a gradual departure from α\alpha = 1 is observed. For TT = 4 K and TT = 6 K, this departure leads to a regime with α\alpha >> 1, while for TT = 8K and 10 K this deviation leads to sublinear or α\alpha << 1 scaling. At high LL values gradually all phonons become diffusive and α\alpha goes to zero.

There are several striking features to point out from Fig. 12. First, prominent contribution of drifting phonons at 4 K leads to an enhanced superlinear scaling with α\alpha >> 1, representing the signature of phonon Poiseuille flow [20] and therefore prominent phonon hydrodynamics which assists in featuring the Knudsen minimum seen in Fig. 10. Here we mention that even for TT = 4 K, the exponent α\alpha gradually starts from 1, reaches a maximum value around LL = 0.8 μ\mum, and goes sublinear with α\alpha << 1 thereafter before going to zero at very high LL values. Therefore, sublinear scaling is universal in the intermediate regime. For TT = 4 K, however, the sublinear scaling precedes a superlinear behavior displaying strong hydrodynamic feature. Second, a minute superlinear scaling, observed in Fig. 11(b) inset for TT = 6 K, can be realized in a better way by observing the small LL-window for which α\alpha >> 1 for TT = 6 K. At TT = 8K and 10 K, though sublinear scaling is observed in the intermediate regime, it decays to zero in different rates. After LL = 10 μ\mum, the decay rate seems faster than that of the cases below LL = 10 μ\mum.

We understand that although average scattering rate and Knudsen number variation with temperature indicates phonon hydrodynamics to present in GeTe for several temperature and characteristic size window, low-κL\kappa_{L} material GeTe needs several factors to manifest a strong hydrodynamic response by phonons. In this context, superlinear scaling of κL\kappa_{L} on LL plays a crucial role in the transition from complete ballistic to complete diffusive propagation regimes. To understand and investigate the reason behind superlinear and sublinear scaling at the intermediate quasi-ballistic regime of phonon transport, we calculate the ratio γ\gamma as a function of LL for three temperatures: TT = 4 K, 6 K and 10 K. We define γ\gamma as

γ=τN1τR1+τB1\gamma=\frac{\left\langle\tau_{N}^{-1}\right\rangle}{\left\langle\tau_{R}^{-1}\right\rangle+\left\langle\tau_{B}^{-1}\right\rangle} (31)

where τN1\left\langle\tau_{N}^{-1}\right\rangle, τR1\left\langle\tau_{R}^{-1}\right\rangle, and τB1\left\langle\tau_{B}^{-1}\right\rangle are average scattering

Refer to caption
Figure 13: (a) LL dependence of γ\gamma for three representative temperatures: TT = 4 K, 6 K and 10 K. The saturation values of γ\gamma (γdiff\gamma_{diff}) in the plateau regimes attained at higher LL values for different temperatures are denoted via blue dashed lines. The red dotted line represents γ\gamma = 1 and the differences between γ\gamma = 1 and γdiff\gamma_{diff} are shown via black double headed arrows. (b) The variation of log(γ)/log(L)\partial log(\gamma)/\partial log(L) as a function of LL for three representative temperatures: TT = 4 K, 6 K and 10 K.

rates for normal, resistive, and boundary scattering respectively. Figure 13.(a) shows that γ\gamma increases gradually and reaches a plateau as we increase LL. In the ballistic phonon conduction regime, we observe γ(T=4K)\gamma(T=4K) << γ(T=6K)\gamma(T=6K) << γ(T=10K)\gamma(T=10K). This is due to the effect of strong boundary scattering at low temperature and low LL. However, in the regime of complete diffusive propagation of phonons a reverse trend is observed: γ(T=4K)\gamma(T=4K) >> γ(T=6K)\gamma(T=6K) >> γ(T=10K)\gamma(T=10K) as in this regime, γ\gamma is independent of size. We define these saturation values as γdiff\gamma_{diff}. Again, the crucial crossover is observed in the intermediate LL-regime. We also mark the difference between γ\gamma = 1 and γdiff\gamma_{diff} in Fig. 13(a) via double headed arrows. This difference characterizes the relative strength of normal scattering compared to the dissipative resistive scattering and therefore indicates the strength for persisting coherent phonon flow.

However, we tend to understand the reason behind the nonlinear behavior of κL\kappa_{L} at the intermediate LL-regime. For this purpose, we present the variation of the exponent of γ\gamma by calculating log(γ)/log(L)\partial log(\gamma)/\partial log(L) as a function of LL in Fig. 13(b). We observe that the exponent for TT = 4 K is higher and for both TT= 4 K and 6 K, it stays around 1 (γ\gamma being linearly increasing with LL) in the intermediate regime. However, for TT = 10 K, log(γ)/log(L)\partial log(\gamma)/\partial log(L) drops up to several orders (at LL = 10 μ\mum, it drops around 10 times) compared to the TT = 4 K and TT = 6 K cases. Therefore, the higher the exponent log(γ)/log(L)\partial log(\gamma)/\partial log(L) in the intermediate regime and closer to 1, the higher the chances of the collective phonon flow due to strong normal scattering. This eventually can lead to the strong appearance of phonon hydrodynamics with superlinear LL dependence of κL\kappa_{L} and prominent Knudsen minimum apart from other signatures born out of the assessment of Knudsen number and average scattering rate as a function of temperature.

V Summary and conclusions

We employ KCM in conjunction with first-principles density functional calculations to investigate the effect of characteristic size (LL) on collective phonon transport in low-thermal conductivity material GeTe. We observe phonon hydrodynamics in crystalline GeTe and identify the competitive effects of both temperature and LL on the collective phonon transport. As a first step, we distinguish heat transport regimes correspond to ballistic and completely diffusive phonon transport. These regimes have been identified as a function of both temperature and LL. In the ballistic regime, the frequency dependence of phonon propagation is understood. Temperature is found to dominate over LL in deciding the excitation of acoustic and optical phonons. Even for very small LL values, correspond to ballistic transport regime, we observe a small contribution coming from optical modes of GeTe if temperature is raised to 30 K. However, at low temperature (TT = 10 K), only acoustic modes excite to enable ballistic propagation. The variation of mean free paths as a function of frequencies also represents this dependence. At low temperatures, increasing LL gradually weakens ballistic conduction. On the other hand, for the same LL value, increasing temperature also gradually weakens the ballistic conduction.

After looking at ballistic and diffusive phonon conduction regimes, we turn our attention toward the intriguing intermediate LL-regime where both ballistic and diffusive phonons are present. The average scattering rates seem to follow the Guyer and Krumhansl hierarchy at low temperatures, indicating the presence of phonon hydrodynamics at certain temperatures and LL window. KCM method allows us to distinguish the variation of collective contribution as functions of both temperature and LL. Therefore, the phonon hydrodynamic regimes in terms of both temperature and LL have been realized using non-local length and Knudsen number (Kn) evaluation which draws a parallel between fluid hydrodynamics and the collective flow of phonons. The hydrodynamic regimes identified using scattering rates are found to satisfy the condition 0.1 << Kn << 10, which is the condition for hydrodynamic flow in terms of Kn. Further, exploiting the variation of lattice thermal conductivity as a function of LL, a heat wave propagation length has been extracted for different temperatures. Comparing this characteristic length scale with phonon propagation lengths reveals that the heat wave propagation length is well captured by phonon propagation with only resistive damping. the Knudsen number can also be associated with this length scale which shows almost similar behavior as that of the Kn obtained using a nonlocal length. For both of these definitions of Kn, the variation of normalized thermal conductivity (κL\kappa_{L}^{*} = κL/L\kappa_{L}/L) with 1/Kn shows a Knudsen minimum like feature only at very low temperature (TT = 4 K). Though Kn can capture the hydrodynamic regimes well in terms of both temperature and LL, some of the prominent features of phonon hydrodynamics, like Knudsen minimum, can be weakly present or may be absent in low-thermal conductivity materials. We have found that the intermediate LL-regime and the scaling of thermal conductivity with LL in this regime works as a marker to determine the existence of the Knudsen minimumlike prominent hydrodynamic feature. A superlinear scaling in this intermediate LL-regime seems to assist a Knudsen minimum and therefore prominent phonon hydrodynamics. On the other hand, sublinear scaling does not lead to a Knudsen like minimum, featuring weak phonon hydrodynamics at those temperatures. A ratio of average normal and resistive scattering rates have been found to control the strength and prominent visibility of the collective phonon transport in GeTe.

In summary, this paper reveals crucial details about how and when the prominent signatures of phonon hydrodynamics can be observed in low-thermal conductivity materials. In this context, it demonstrates and systematically analyzes the consequences of the competitive effects between temperature and characteristic size on phonon hydrodynamics in GeTe. The outcome of this study can lead to a better understanding of the generic behavior and consequences of the phonon hydrodynamics and its controlling parameters in any other low-thermal conductivity materials. The accurate description of phonon hydrodynamics in low-κ\kappa materials can also lead to better theoretical predictions of experimentally observed thermal conductivity at low temperatures for these materials.

Acknowledgements.
This project has received funding from the European Union’s Horizon 2020 research and innovation program under Grant Agreement No. 824957 (“BeforeHand:” Boosting Performance of Phase Change Devices by Hetero- and Nanostructure Material Design).

VI Supplementary Material

VI.1 Collective and Kinetic mean free path

Figure S1 presents the variation of effective mean free paths of GeTe as a function of phonon frequency for two different LL values at TT = 10 K. In KCM approach, mean free paths can be distinguished as kinetic and collective mean free paths. As described in the main text, the kinetic mean free paths (lk(ω)l_{k}(\omega)) are different for different modes but the collective mean free paths (lc(T)l_{c}(T)) are same for all modes and it is only a function of temperature. The effective mean free path has been realized using leff(ω)l_{eff}(\omega) = (1-Σ\Sigma)lk(ω)l_{k}(\omega)+Σ\Sigma lcl_{c}. In Fig S1.(a), we observe that at TT = 10 K, the dominating contribution comes from the kinetic mean free path for LL = 0.003 μ\mum. We recall from the main text that this LL corresponds to ballistic phonon conduction regime for GeTe at 10 K. However, LL = 0.8 μ\mum satisfies the Guyer and Krumhansl condition for phonon hydrodynamics and consistently the collective mean free path is seen to dominate over the kinetic mean free path (Fig S1.(b)).

Refer to caption
Figure S1: Effective mean free path (MFP) of crystalline GeTe, along with collective and kinetic contributions are presented as a function of frequencies for two different LL at TT = 10 K: (a) LL = 0.003 μ\mum and (b) LL = 0.8 μ\mum. Gray shaded regions denote acoustic mode frequency regime for GeTe.

VI.2 Variation of Knudsen number with temperature for different LL using heat wave propagation length

Figure S2 presents the variation of Knudsen number (Kn) as a function of temperature for different LL values. Kn has been realized via the heat wave propagation length (LhwplL_{hwpl}) as LhwplL_{hwpl}/LL, where LhwplL_{hwpl} is obtained as a characteristic length at which the lattice thermal conductivity in the completely diffusive limit correspond to bulk sample reduces to 1/e times. The hydrodynamic regime follows 0.1 << Kn << 10 and therefore has been marked with the shaded region. The hydrodynamic LL-regimes obtained using average scattering rates are also shown via rectangular boxes for TT = 6 K, 10 K and 20 K. Except the ballistic hydrodynamic boundary regimes for TT = 6 K, the regimes evaluated by Kn and average scattering rates are found to be consistent. The transition between ballistic and hydrodynamic regimes are often found to be blurry and without sharp demarcation in low-thermal conductivity materials. This has been discussed in the main text.

Refer to caption
Figure S2: The variation of Knudsen number (Kn) with temperature for different LL values of crystalline GeTe. The shaded region satisfies 0.1 \leq Kn \leq 10 while the rectangular boxes define phonon hydrodynamic regimes calculated from average scattering rates. Blue dashed lines to guide the eye for TT = 6 K, 10 K and 20 K.

References

  • Cepellotti et al. [2015] A. Cepellotti, G. Fugallo, L. Paulatto, M. Lazzeri, F. Mauri, and N. Marzari, Nat Commun 6, 6400 (2015).
  • Gill-Comeau and Lewis [2015] M. Gill-Comeau and L. J. Lewis, Appl. Phys. Lett. 106, 193104 (2015).
  • Li and Lee [2018] X. Li and S. Lee, Phys. Rev. B 97, 094309 (2018).
  • Lee and Li [2020] S. Lee and X. Li, Hydrodynamic phonon transport: past, present and prospects, in Nanoscale Energy Transport, 2053-2563 (IOP Publishing, Bristol, UK, 2020) pp. 1–1 to 1–26.
  • Lindsay et al. [2019] L. Lindsay, A. Katre, A. Cepellotti, and N. Mingo, Journal of Applied Physics 126, 050902 (2019).
  • Hardy [1970] R. J. Hardy, Phys. Rev. B 2, 1193 (1970).
  • Guo and Wang [2015] Y. Guo and M. Wang, Physics Reports 595, 1 (2015).
  • Machida et al. [2018] Y. Machida, A. Subedi, K. Akiba, A. Miyake, M. Tokunaga, Y. Akahama, K. Izawa, and K. Behnia, Science Advances 410.1126/sciadv.aat3374 (2018).
  • Markov et al. [2018] M. Markov, J. Sjakste, G. Barbarino, G. Fugallo, L. Paulatto, M. Lazzeri, F. Mauri, and N. Vast, Phys. Rev. Lett. 120, 075901 (2018).
  • Cepellotti and Marzari [2017] A. Cepellotti and N. Marzari, Phys. Rev. Materials 1, 045406 (2017).
  • Guyer and Krumhansl [1966a] R. A. Guyer and J. A. Krumhansl, Phys. Rev. 148, 766 (1966a).
  • Guyer and Krumhansl [1966b] R. A. Guyer and J. A. Krumhansl, Phys. Rev. 148, 778 (1966b).
  • Fugallo et al. [2014] G. Fugallo, A. Cepellotti, L. Paulatto, M. Lazzeri, N. Marzari, and F. Mauri, Nano Letters 14, 6109 (2014).
  • Cepellotti and Marzari [2016] A. Cepellotti and N. Marzari, Phys. Rev. X 6, 041013 (2016).
  • Sendra et al. [2021] L. Sendra, A. Beardo, P. Torres, J. Bafaluy, F. X. Alvarez, and J. Camacho, Phys. Rev. B 103, L140301 (2021).
  • Li and Lee [2019] X. Li and S. Lee, Phys. Rev. B 99, 085202 (2019).
  • Ho et al. [2018] D. Y. H. Ho, I. Yudhistira, N. Chakraborty, and S. Adam, Phys. Rev. B 97, 121404 (2018).
  • Machida et al. [2020] Y. Machida, N. Matsumoto, T. Isono, and K. Behnia, Science 367, 309 (2020).
  • Huberman et al. [2019] S. Huberman, R. A. Duncan, K. Chen, B. Song, V. Chiloyan, Z. Ding, A. A. Maznev, G. Chen, and K. A. Nelson, Science 364, 375 (2019).
  • Ding et al. [2018a] Z. Ding, J. Zhou, B. Song, V. Chiloyan, M. Li, T.-H. Liu, and G. Chen, Nano Letters 18, 638 (2018a).
  • Narayanamurti and Dynes [1972] V. Narayanamurti and R. C. Dynes, Phys. Rev. Lett. 28, 1461 (1972).
  • Koreeda et al. [2007] A. Koreeda, R. Takano, and S. Saikan, Phys. Rev. Lett. 99, 265502 (2007).
  • Martelli et al. [2018] V. Martelli, J. L. Jiménez, M. Continentino, E. Baggio-Saitovitch, and K. Behnia, Phys. Rev. Lett. 120, 125901 (2018).
  • Koreeda et al. [2010] A. Koreeda, R. Takano, A. Ushio, and S. Saikan, Phys. Rev. B 82, 125103 (2010).
  • Ackerman et al. [1966] C. C. Ackerman, B. Bertman, H. A. Fairbank, and R. A. Guyer, Phys. Rev. Lett. 16, 789 (1966).
  • Jackson et al. [1970] H. E. Jackson, C. T. Walker, and T. F. McNelly, Phys. Rev. Lett. 25, 26 (1970).
  • Torres et al. [2019] P. Torres, F. X. Alvarez, X. Cartoixà, and R. Rurali, 2D Materials 6, 035002 (2019).
  • Ghosh et al. [2020a] K. Ghosh, A. Kusiak, and J.-L. Battaglia, Phys. Rev. B 102, 094311 (2020a).
  • Levin et al. [2013] E. M. Levin, M. F. Besser, and R. Hanus, J. Appl. Phys. 114, 083713 (2013).
  • Campi et al. [2015] D. Campi, D. Donadio, G. C. Sosso, J. Behler, and M. Bernasconi, J. Appl. Phys. 117, 015304 (2015).
  • Torres et al. [2017] P. Torres, A. Torelló, J. Bafaluy, J. Camacho, X. Cartoixà, and F. X. Alvarez, Phys. Rev. B 95, 165407 (2017).
  • Ghosh et al. [2020b] K. Ghosh, A. Kusiak, P. Noé, M.-C. Cyrille, and J.-L. Battaglia, Phys. Rev. B 101, 214305 (2020b).
  • Togo et al. [2015] A. Togo, L. Chaput, and I. Tanaka, Phys. Rev. B 91, 094306 (2015).
  • Giannozzi et al. [2009] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. D. Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, J. Phys.: Condens.Matter 21, 395502 (2009).
  • Monkhorst and Pack [1976] H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).
  • Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77, 3865 (1996).
  • Shaltaf et al. [2009] R. Shaltaf, X. Gonze, M. Cardona, R. K. Kremer, and G. Siegle, Phys. Rev. B 79, 075204 (2009).
  • Campi et al. [2017] D. Campi, L. Paulatto, G. Fugallo, F. Mauri, and M. Bernasconi, Phys. Rev. B 95, 024311 (2017).
  • Blochl [1994] P. E. Blochl, Phys. Rev. B 50, 17953 (1994).
  • Mizokami et al. [2018] K. Mizokami, A. Togo, and I. Tanaka, Phys. Rev. B 97, 224306 (2018).
  • ichiro Tamura [1983] S. ichiro Tamura, Phys. Rev. B 27, 858 (1983).
  • Laeter et al. [2003] J. R. D. Laeter, J. K. Böhlke, P. D. Bièvre, H. Hidaka, H. S. Peiser, K. J. R. Rosman, and P. D. P. Taylor, Pure Appl. Chem. 75, 683 (2003).
  • Kaviany [2014] M. Kaviany, Heat Transfer Physics, 2nd ed. (Cambridge University Press, New York, NY, 2014).
  • Alvarez [2018] P. Alvarez, Thermal Transport in Semiconductors: First Principles and Phonon Hydrodynamics, Springer Theses (Springer International Publishing, Cham, Switzerland, 2018).
  • Saito et al. [2018] R. Saito, M. Mizuno, and M. S. Dresselhaus, Phys. Rev. Applied 9, 024017 (2018).
  • Park et al. [2013] M. Park, S.-C. Lee, and Y.-S. Kim, J. Appl. Phys. 114, 053506 (2013).
  • Xu et al. [2014] X. Xu, L. F. Pereira, Y. Wang, J. Wu, K. Zhang, X. Zhao, S. Bae, C. T. Bui, R. Xie, J. T. Thong, B. H. Hong, K. P. Loh, D. Donadio, B. Li, and B. Ozyilmaz, Nat Commun 5, 3689 (2014).
  • Ding et al. [2018b] Z. Ding, J. Zhou, B. Song, M. Li, T.-H. Liu, and G. Chen, Phys. Rev. B 98, 180302 (2018b).