Phonon BTE

Introduction to the BTE

This section follows the theory of the reference available at this link (https://arxiv.org/abs/1212.0470). Here, we report the most relevant points of the manuscript.

When a gradient of temperature \(\nabla T\) is applied to a system, a subsequent heat flux will start propagating in the medium. Without loss of generality, we assume the gradient of temperature to be along the direction \(x\).

The flux of heat, collinear to the temperature gradient, can be written in terms of the phonon energies \(\hbar\omega_{\boldsymbol{q}j}\), phonon group velocities \(v_{q j}\) in the \(x\) direction, and the perturbed phonon population \(n_{q j}\),

\[\frac{1}{N_0 \Omega} \sum_{q j} \hbar \omega_{q j} \boldsymbol{v}_{\boldsymbol{q} j} n_{\boldsymbol{q} j} = - k \frac{\partial T}{ \partial \boldsymbol{x}}\]

On the l.h.s., \(\omega_{\boldsymbol{q}j }\) is the angular frequency of the phonon mode with vector \(\boldsymbol{q}\) and branch index \(j\), \(\Omega\) is the volume of the unit cell, and the summation runs over a uniform mesh of \(N_0 \boldsymbol{q}\) points. On the r.h.s., \(k\) is the diagonal component of the thermal conductivity in the temperature-gradient direction. We will use \(\alpha\) and \(\beta\) labeling the Cartesian indices.

From this expression, we can see that knowledge of the perturbed phonon population allows heat flux and subsequently thermal conductivity to be evaluated. Unlike phonon scattering by defects, impurities and boundaries, anharmonic scattering represents an intrinsic resistive process and in high quality samples, at room temperature, it dominates the behaviour of lattice thermal conductivity balancing the perturbation due to the gradient of temperature. The balance equation, namely the Boltzmann Transport Equation (BTE), formulated in 1929 by Peierls is,

\[-\boldsymbol{v}_{q j}\cdot \frac {\partial T} {\partial \boldsymbol{x}} \frac{\partial n_{\boldsymbol{q} j}}{\partial T} + \frac{\partial n_{\boldsymbol{q} j}}{\partial t}\bigg|_{scatt} = 0,\]

with the first term indicating the phonon diffusion due to the temperature gradient and the second term the scattering rate due to all the scattering processes. This equation has to be solved self consistently. In the general approach, for small perturbation from the equilibrium, the temperature gradient of the perturbed phonon population is replaced with the temperature gradient of the equilibrium phonon population \(\partial n_{\boldsymbol{q} j} / \partial T = \partial \bar{n}_{\boldsymbol{q} j} / \partial T\) where \(\bar{n}_{\boldsymbol{q} j} = (e^{\hbar \omega_{\boldsymbol{q} j} /k_BT} - 1)^{-1}\), while the scattering term can be expanded about its equilibrium value in terms of a first-order perturbation, \(f^{\mathrm{EX}}\),

\[n_{\boldsymbol{q} j} \simeq \bar{n}_{\boldsymbol{q} j}+\bar{n}_{\boldsymbol{q} j}(\bar{n}_{\boldsymbol{q} j}+1) \frac{\partial T}{\partial \boldsymbol{x}}\cdot f^{\mathrm{EX}}_{\boldsymbol{q} j}\]

The linearized BTE can then be written in the following form:

\[\begin{split}-v_{\boldsymbol{q} j}\left(\frac{\partial \bar{n}_{\boldsymbol{q} j}}{\partial T}\right) = \sum_{\boldsymbol{q}' j',\boldsymbol{q}'' j''}\Big[ P_{\boldsymbol{q} j,\boldsymbol{q}' j'}^{\boldsymbol{q}'' j''}(f^{\mathrm{EX}}_{\boldsymbol{q} j}+f^{\mathrm{EX}}_{\boldsymbol{q}' j'}-f^{\mathrm{EX}}_{\boldsymbol{q}'' j''}) + \frac{1}{2} P^{\boldsymbol{q}' j',\boldsymbol{q}'' j''}_{\boldsymbol{q} j} (f^{\mathrm{EX}}_{\boldsymbol{q} j}-f^{\mathrm{EX}}_{\boldsymbol{q}' j'}-f^{\mathrm{EX}}_{\boldsymbol{q}'' j''} )\Big] \\\\ + \sum_{\boldsymbol{q}' j'} P^{\mathrm{isot}}_{\boldsymbol{q} j,\boldsymbol{q}' j'} (f^{\mathrm{EX}}_{\boldsymbol{q} j} - f^{\mathrm{EX}}_{\boldsymbol{q}' j'}) + P^{\mathrm{be}}_{\boldsymbol{q} j} f^{\mathrm{EX}}_{\boldsymbol{q} j}\end{split}\]

where the sum over \(\boldsymbol{q}'\) and \(\boldsymbol{q}^{''}\) is performed over the first Brillouin zone. The \(\mathrm{EX}\) superscript of the first-order perturbation \(f^{\mathrm{EX}}\) denotes the exact solution of the BTE, to distinguish it from the various approximate solutions to be discussed later.

Phonon scattering rates

In this last equation, anharmonic scattering processes, as well as the scattering with isotopic impurities and boundaries, are included in place of the scattering term. We define these rates below.

Anharmonic phonon scattering

In order to evaluate these anharmonic (3-phonon) scattering rates, it is necessary to compute the third derivative, \(V^{(3)}\), of the total energy of the crystal \(\mathcal{E}^{tot}(\{u_{s \alpha} (\boldsymbol{R}_l) \})\) with respect to the atomic displacement \(u_{s \alpha} (\boldsymbol{R}_l)\) from the equilibrium position of the s-th atom along the Cartesian coordinate \(\alpha\) in the crystal cell identified by the lattice vector \(\boldsymbol{R}_l\),

\[V^{(3)}(\boldsymbol{q} j,\boldsymbol{q}' j',\boldsymbol{q}'' j'')= \frac{\partial^3 \mathcal{E}^{cell}} {\partial X_{\boldsymbol{q} j},\partial X_{\boldsymbol{q}' j'},\partial X_{\boldsymbol{q}'' j''}}\]

where \(\mathcal{E}^{cell}\) is the energy per unit cell. The non-dimensional quantity \(X_{\boldsymbol{q} j}\) is defined by

\[X_{\boldsymbol{q} j}= \frac{1}{N_0}\sum_{l,s,\alpha} \sqrt{\frac{2 M_s \omega_{\boldsymbol{q} j}} {\hbar}} z^{s \alpha^*}_{\boldsymbol{q} j} u_{s \alpha }(\boldsymbol{R}_l) e^{-i\boldsymbol{q}\cdot \boldsymbol{R}_l}\]

where \(z^{s \alpha}_{\boldsymbol{q}j}\) is the orthogonal phonon eigenmode normalized on the unit cell and \(M_s\) the atomic masses.

This expression for \(X\) can be used to transform the output of a density functional theory code, i.e. the matrix of energy derivatives in real space \(\mathcal{E}(\boldsymbol{R}_l s\alpha,\boldsymbol{R}'_{l'} s' \alpha',\boldsymbol{R}''_{l''}s''\alpha'')\) to the Fourier space.

The matrix is actually a periodic function, so it is possible to neglect one of the Bravais lattice vector indices of such a tensor. Note that Quantum Espresso provides the matrix \(\mathcal{E}(\boldsymbol{0} s\alpha,\boldsymbol{R}'_{l'} s' \alpha',\boldsymbol{R}''_{l''}s''\alpha'')\), where they used this freedom to set the first vector to the initial unit cell. Codes like phono3py instead provide the complete matrix, and we select the subset \(\mathcal{E}(\boldsymbol{0} s\alpha,\boldsymbol{R}'_{l'} s' \alpha',\boldsymbol{R}''_{l''}s''\alpha'')\) that we will later use.

The Fourier transform of these coupling matrix elements is,

\[V^{(3)}(\boldsymbol{q}s\alpha,\boldsymbol{q}'s'\alpha',\boldsymbol{q}''s''\alpha'') = \sum_{\boldsymbol{R}_l, \boldsymbol{R}_{l'}} \mathcal{E}(\boldsymbol{R}_{l} s\alpha,\boldsymbol{R}'_{l'} s' \alpha',0 s''\alpha'')\]

where it’s worth noting that the sum over Bravais lattice vectors is done over the Bravais lattice vectors \(\boldsymbol{R}_l\) such that \(\boldsymbol{R}_l\) belongs to the Wigner Seitz zone of the super cell in which the anharmonic force constants have been computed.

These coupling matrix elements can then be used to calculate anharmonic phonon scattering rates.

More specifically, \(P_{\boldsymbol{q} j,\boldsymbol{q}' j'}^{\boldsymbol{q}'' j''}\) is the equilibrium scattering rate for a process where a phonon mode \(\boldsymbol{q}j\) scatters by absorbing another mode, \(\boldsymbol{q}' j'\) to generate a third phonon mode \(\boldsymbol{q}'' j''\). \(P^{\boldsymbol{q}' j',\boldsymbol{q}'' j''}_{\boldsymbol{q} j}\) is the equilibrium scattering rate of the opposite process, in which a phonon mode \(\boldsymbol{q}j\) decays into two modes, \(\boldsymbol{q}'j'\) and \(\boldsymbol{q}''j''\).

The two scattering rates have the forms:

\[P^{\boldsymbol{q}'' j''}_{\boldsymbol{q} j,\boldsymbol{q}' j'} = \frac{2 \pi}{N_0 \hbar^2} \sum_{\boldsymbol{G}} |V^{(3)}(\boldsymbol{q} j,\boldsymbol{q}' j',-\boldsymbol{q}'' j'')|^2 \bar{n}_{\boldsymbol{q} j}\bar{n}_{\boldsymbol{q}' j'}(\bar{n}_{\boldsymbol{q}'' j''}+1) \delta_{\boldsymbol{q}+\boldsymbol{q}' -\boldsymbol{q}'', \boldsymbol{G}} \delta(\hbar \omega_{\boldsymbol{q} j} +\hbar \omega_{\boldsymbol{q}' j'}-\hbar \omega_{\boldsymbol{q}'' j''})\]
\[P^{\boldsymbol{q}' j',\boldsymbol{q}'' j''}_{\boldsymbol{q} j} = \frac{2 \pi}{N_0 \hbar^2 } \sum_{\boldsymbol{G}} |V^{(3)}(\boldsymbol{q} j,-\boldsymbol{q}' j',-\boldsymbol{q}'' j'')|^2 \bar{n}_{\boldsymbol{q} j}(\bar{n}_{\boldsymbol{q}' j'}+1)(\bar{n}_{\boldsymbol{q}'' j''}+1)\delta_{\boldsymbol{q}-\boldsymbol{q}' -\boldsymbol{q}'', \boldsymbol{G}} \delta(\hbar \omega_{\boldsymbol{q} j}-\hbar \omega_{\boldsymbol{q}' j'}-\hbar \omega_{\boldsymbol{q}'' j''} )\]

where \(\boldsymbol{G}\) are the reciprocal lattice vectors.

Phonon-isotope scattering

The rate of the elastic scattering with isotopic impurities has the form:

\[P_{\boldsymbol{q} j,\boldsymbol{q}' j'}^{\mathrm{isot}} = \frac{\pi}{2 N_0} \omega_{\boldsymbol{q} j}\omega_{\boldsymbol{q}' j'} \left[ \bar{n}_{\boldsymbol{q} j} \bar{n}_{\boldsymbol{q}' j'} + \frac{\bar{n}_{\boldsymbol{q} j} + \bar{n}_{\boldsymbol{q}' j'}} {2} \right ] \sum_{s} g^{s}_{2} \bigg| \sum_{\alpha} z^{s \alpha^*}_{\boldsymbol{q}j} \cdot z^{s \alpha}_{\boldsymbol{q}' j'} \bigg|^2 \delta (\omega_{\boldsymbol{q} j}- \omega_{\boldsymbol{q}' j'})\]

where we introduced the isotopic scattering strength \(g^s_2\):. See https://doi.org/10.1103/PhysRevLett.106.045901 for more details.

\[g^s_2 = \frac{1}{\langle m_s \rangle^2} \sum_i f_{is} (m_{is} - \langle m_s \rangle)^2\]

where \(\big< m_s \big>\) is the average mass of an atom of species \(s\), \(m_{is}\) is the mass of the isotope \(i\) of atomic type \(s\), and \(f_{is}\) is the isotopic abundance.

Phonon-boundary scattering

In a system of finite size, \(P_{q j}^{\mathrm{be}}\) describes the reflection of a phonon from the border,

\[P_{\boldsymbol{q} j}^{\mathrm{be}} = \frac{v_{\boldsymbol{q} j}}{L}\bar{n}_{\boldsymbol{q} j}(\bar{n}_{\boldsymbol{q} j}+1),\]

where \(L\) is the Casimir length of the sample. This boundary scattering is treated in the relaxation time approximation, and it results in a process in which a phonon from a specific state (\(\boldsymbol{q} j\)) is reemitted from the surface, contributing only to the equilibrium distribution.

Solutions of the phonon BTE

Now that we have the linearized BTE and the scattering rates which appear from the scattering term, we can work on solving the BTE. For the sake of clarity, we will contract from here on the vector \(\boldsymbol{q}\) and branch index \(j\) in a single mode index \(\nu\).

The linerized BTE as defined earlier can be written as a linear system in matrix form:

\[\boldsymbol{A} \boldsymbol{f}^{\mathrm{EX}}=\boldsymbol{b}\]

with the vector \(b_{\nu'} =-v_{\nu'}\hbar \omega_{\nu'} \bar{n}_{\nu'}(\bar{n}_{\nu'}+1)\) and the matrix

\[A_{\nu,\nu'} = \left[{\sum_{\nu'',\nu'''}} (P^{\nu''}_{\nu,\nu'''} + \frac{ P_{\nu''',\nu''}^{\nu}}{2} ) + \sum_{\nu''} P^{\mathrm{isot}}_{\nu,\nu''} + P^{\mathrm{be}}_{\nu} \right] \delta_{\nu,\nu'} - {\sum_{\nu''}} ( P^{\nu'}_{\nu,\nu''} -P^{\nu''}_{\nu,\nu'}+ P_{\nu',\nu''}^{\nu} ) + P^{\mathrm{isot}}_{\nu,\nu'}\]

where we have used \(P^{\nu', \nu''}_{\nu}=P_{\nu', \nu''}^{\nu}\) from the detailed balance condition \(\bar{n}_{\nu}(\bar{n}_{\nu'}+1)(\bar{n}_{\nu''}+1) = (\bar{n}_{\nu}+1)\bar{n}_{\nu'}\bar{n}_{\nu''}\) (valid under the assumption \(\hbar \omega = \hbar \omega' + \hbar \omega''\)). In this form the matrix is symmetric and positive semi-definite and it can be decomposed in \(\boldsymbol{A} = \boldsymbol{A}^{\mathrm{out}} +\boldsymbol{A}^{\mathrm{in}}\), where

\[A^{\mathrm{out}}_{\nu,\nu'} = \frac{\bar{n}_{\nu}(\bar{n}_{\nu} +1)} {\tau^{\mathrm{T}}_{\nu}}\delta_{\nu,\nu'}\]
\[A^{\mathrm{in}}_{\nu,\nu'} = - \sum_{\nu''} \left( P^{\nu'}_{\nu,\nu''} -P^{\nu''}_{\nu,\nu'}+ P_{\nu',\nu''}^{\nu} \right ) + P^{\mathrm{isot}}_{\nu,\nu'}\]

with \(\tau^{\mathrm{T}}_{\nu}\) being the phonon relaxation time. The \(\boldsymbol{A}^{\mathrm{out}}\) diagonal matrix describes the depopulation of phonon states due to the scattering mechanisms while the \(\boldsymbol{A}^{\mathrm{in}}\) matrix describes their repopulation due to the incoming scattered phonons.

The solution of the linear system is obtained formally by inverting the matrix \({\boldsymbol A}\),

\[{\boldsymbol f}^{\mathrm{EX}} = \frac{1}{\boldsymbol{A}} {\boldsymbol b}\]

and subsequently, the thermal conductivity will be evaluated as:

\[k = \lambda {\boldsymbol b} \cdot {\boldsymbol f}^{\mathrm{EX}} = - \frac{\hbar}{N_0\Omega k_B T^2}\sum_{\nu}v_{\nu} \omega_{\nu} \bar{n}_{\nu}(\bar{n}_{\nu}+1) f_{\nu}^{\mathrm{EX}}\]

with \(\lambda= 1 /(N_0\Omega k_B T^2)\).

RTA solution

In the relaxation time approximation (RTA), we set \(\boldsymbol{A}^{\mathrm{in}}\) to zero so that,

\[{\boldsymbol f}^{\mathrm{SMA}} =\frac{1}{ \boldsymbol{A}^{\mathrm{out}}} {\boldsymbol b}.\]

Inverting \(\boldsymbol{A}^{\mathrm{out}}\) is trivial due to its diagonal form. The lattice thermal conductivity in RTA is then

\[k^{\mathrm{RTA}}=\lambda \boldsymbol{b} \cdot \boldsymbol{f}^{\mathrm{SMA}}=\frac{\hbar^2}{N_0\Omega k_B T^2}\sum_{\nu}v^2_{\nu} \omega^2_{\nu} \bar{n}_{\nu}(\bar{n}_{\nu}+1)\tau^{\mathrm{T}}_{\nu}.\]

Iterative solution: Omini-Sparavigna method

Note

Generally, we recommend the variational method over this, as the variational method converges more quickly. However, this method can be computationally cheaper in Phoebe, as it enables one to take advantage of symmetries.

An exact solution of the BTE that does not imply either storing or the explicit inversion of matrix \(\boldsymbol{A}\) has been proposed by Omini and Sparavigna by converging with respect to the iteration \(i\) the following:

\[\boldsymbol{f}_{ i+1} =\frac{1} {\boldsymbol{A}^{\mathrm{out} } } \boldsymbol{b} - \frac{1} {\boldsymbol{A}^{\mathrm{out} } } \boldsymbol{A}^{\mathrm{in}} \boldsymbol{f}_{i}\]

with the iteration zero consisting in the RTA \(\boldsymbol{f}_0=\boldsymbol{f}^{\mathrm{RTA}}\). Instead of storing and inverting \(\boldsymbol{A}\), it just requires the evaluation of \(\boldsymbol{A}^{\mathrm{in}}\:\boldsymbol{f}_{i}\), at each iteration \(i\) of the OS method, which is an operation computationally much less demanding. Once the convergence is obtained the thermal conductivity is evaluated by:

\[k^{\mathrm{NV}}(\boldsymbol{f}_i)=\lambda \boldsymbol{b}\cdot \boldsymbol{f}_{i}\]

From a mathematical point of views the OS iterative procedure can be written as a geometric series:

\[\boldsymbol{f}_{ i} = \sum_{j=0,i} \left(-\frac{1}{\boldsymbol{A}^{\mathrm{out}}} \boldsymbol{A}^{\mathrm{in}}\right)^{j} \frac{1}{\boldsymbol{A}^{\mathrm{out}}} \: \boldsymbol{b} \;.\]

Iterative solution: Variational method

An alternative approach consists in using the properties of the matrix \({\boldsymbol A}\) to find the exact solution of the linearized BTE, via the variational principle. The solution of the BTE is the vector \(\boldsymbol{f}^{\mathrm{EX}}\) which makes stationary the quadratic form

\[\mathcal{F}(\boldsymbol{f}) =\frac{1}{2} {\boldsymbol f} \cdot{\boldsymbol A} {\boldsymbol f}- {\boldsymbol b} \cdot {\boldsymbol f}\]

for a generic vector \(\boldsymbol{f}\). Since \(\boldsymbol{A}\) is positive, the stationary point is the global and single minimum of this functional. One can then define a variational conductivity functional:

\[k^\mathrm{V}(\boldsymbol{f}) = - 2 \lambda \mathcal{F}({\boldsymbol f})\]

that has the property \(k^\mathrm{V}(\boldsymbol{f}^{\mathrm{EX}})=k\) while any other value of \(k^{\mathrm{V}}(\boldsymbol{f})\) underestimates \(k\). In other words, finding the minimum of the quadratic form is equivalent to maximizing the thermal conductivity functional. As a consequence, an error in \(f\), \(\delta \boldsymbol{f}= \boldsymbol{f} - \boldsymbol{f}^{\mathrm{EX}}\), results in an error in conductivity, linear in \(\delta \boldsymbol{f}\) when using the non-variational estimator, and quadratic in the variational form.

Here we solve the BTE on a grid (as in OS procedure) but now using the conjugate gradient method, to obtain the exact solution of the BTE equation. In order to speed up the convergence of the conjugate gradient we take advantage of the diagonal and dominant role of \(\boldsymbol{A}^{\mathrm{out}}\) and we use a preconditioned conjugate gradient. Formally, this corresponds using the rescaled variable,

\[\tilde{{\boldsymbol f}} = \sqrt{{\boldsymbol A^{\mathrm{out}}}} {\boldsymbol f}\]

as a preconditioner. Then, with respect to this new variable, we minimize the quadratic form \(\tilde{\mathcal{F}}(\tilde{\boldsymbol{f}}) = \mathcal{F}(\boldsymbol{f})\) where:

\[\tilde{\mathcal{F}}( \tilde{\boldsymbol{f}}) =\frac{1}{2} \tilde{\boldsymbol{f}}\cdot \tilde{\boldsymbol{A}} \tilde{\boldsymbol{f}}- \tilde{\boldsymbol{ b}}\cdot\tilde{\boldsymbol {f}}\]

and

\[\tilde{{\boldsymbol A}} =\frac{1}{ \sqrt{{\boldsymbol A^{\mathrm{out}}}}} {\boldsymbol A}\frac{1}{ \sqrt{{\boldsymbol A^{\mathrm{out}}}}}\]
\[\tilde{{\boldsymbol b}} =\frac{1}{ \sqrt{{\boldsymbol A^{\mathrm{out}}}}} {\boldsymbol b} \label{prec3}\]

Notice that \(\tilde{\boldsymbol{f}}^{\mathrm{RTA}}=\tilde{\boldsymbol{b}}\). The square root evaluation of \(\boldsymbol{A}^{\mathrm{out}}\) is trivial due to its diagonal form. The computational cost per iteration of the conjugate gradient scheme is equivalent to the OS one, but it should have much better convergence and requires a smaller number of iterations.

The conjugate gradient minimization requires the evaluation of the gradient \(\boldsymbol{g}_i= \boldsymbol{A} \boldsymbol{f}_i - \boldsymbol{b}\) and a line minimization. Since the form is quadratic, the line minimization can be done analytically and exactly. Moreover the information required by the line minimization at iteration \(i\) can be recycled to compute the gradient at the next iteration \(i+1\). Starting with an the initial vector \(\boldsymbol{f}_0= \boldsymbol{f}^{\mathrm{RTA}}\), initial gradient \(\boldsymbol{g}_0=\boldsymbol{A}\boldsymbol{f}_0 -\boldsymbol{f}^{\mathrm{RTA}}\) and letting \(\boldsymbol{h}_0= -\boldsymbol{g}_0\), the conjugate gradient method can be summarized with the recurrence:

\[\boldsymbol{t}_i =\boldsymbol{A} \boldsymbol{h}_i\]
\[{\boldsymbol f}_{i+1} = {\boldsymbol f}_{i} - \frac{\boldsymbol {g}_{i} \cdot {\boldsymbol{h}_{i}} } {\boldsymbol{h}_{i} \cdot \boldsymbol{t}_i } \boldsymbol{h}_{i}\]
\[\boldsymbol{g}_{i+1} = \boldsymbol{g}_{i}-\frac{\boldsymbol {g}_{i} \cdot {\boldsymbol{h}_{i}} } {\boldsymbol{h}_{i} \cdot \boldsymbol{t}_i }\boldsymbol{t}_i\]
\[\boldsymbol{h}_{i+1} = -\boldsymbol{g}_{i+1} + \frac{\boldsymbol{g}_{i+1} \cdot \boldsymbol{g}_{i+1}}{{\boldsymbol{g}_{i}} \cdot {\boldsymbol{g}_{i}} } {\boldsymbol h}_{i}\]

where \(\boldsymbol{h}_i\) is the search direction and \(\boldsymbol{t}_i\) is an auxiliary vector. Notice that each iteration requires only one application of the matrix \(\boldsymbol{A}\) on the vector \(\boldsymbol{h}_i\) as in the OS method.

Relaxons solution

In the relaxons method, (presented here Physical Review X 6, no. 4 (2016): 041013.), we first directly diagonalize the scattering matrix:

\[\frac{1}{N_k} \sum_{\nu} \Omega_{\nu\nu'} \theta_{\nu'\alpha} = \frac{1}{\tau_{\alpha}} \theta_{\nu\alpha}\]

where \(\theta\) are relaxons eigenvectors, \(\alpha\) are eigenvalue indices, \(\frac{1}{\tau_{\alpha}}\) are eigenvalues, and the scattering matrix is:

\[\Omega_{\nu\nu'} = \frac{ A_{\nu\nu'} } { \sqrt{ \bar{n}_{\nu}(\bar{n}_{\nu}+1) \bar{n}_{\nu'}(\bar{n}_{\nu'}+1) } }\]

Next, we compute the velocities:

\[\boldsymbol{V}_{\alpha} = \frac{1}{N_k} \sum_{\nu} \theta_{\nu0} \boldsymbol{v}_{\nu} \theta_{\nu\alpha}\]

where

\[\theta_{\nu0} = \sqrt { \frac{ \frac{\partial \bar{n}_{\nu}}{\partial \epsilon} }{C T} } \hbar \omega_{\nu}\]

Finally, the thermal conductivity is:

\[k^{ij} = \sum_{\alpha} C V_{\alpha}^i V_{\alpha}^j \tau_{\alpha}\]

Phonon Velocity Operator

The velocity operator matrix elements (e.g. along the x direction) can be computed using the Hellmann-Feynman theorem from the Dynamical matrix \(\boldsymbol{\mathcal{D}}\):

\[V^x_{j j'}(\boldsymbol{q}) = \sum_{\alpha \alpha' s s'} \frac{1}{2 \sqrt{M_s M_{s'}} \omega_{\boldsymbol{q} j} } z^{s \alpha^*}_{\boldsymbol{q} j} \frac{\partial \mathcal{D}^{\alpha \alpha'}_{s s'}(\boldsymbol{q})}{ \partial q_x} z^{s' \alpha'}_{\boldsymbol{q} j'}\]

In the case of non-degenerate phonon modes, the group velocity is \(\boldsymbol{v}_{\boldsymbol{q} j}=\boldsymbol{V}_{j j}(\boldsymbol{q})\) while when degenerate modes are present, we use the phonon polarization vectors that diagonalize the matrix in the degenerate subspace.

Wigner correction to phonon thermal conductivity

The Wigner transport equation theory is fully described in the reference available at this link (https://www.nature.com/articles/s41567-019-0520-x).

Extremely briefly, thermal conductivity including the Wigner correction is estimated as:

\[k_{\alpha\beta} = k^{BTE}_{\alpha\beta} + \frac{k_BT^2}{\Omega N_k} \sum_{\boldsymbol{q}} \sum_{s\neq s'} \frac{\omega_{\boldsymbol{q}j}+\omega_{\boldsymbol{q}j'}}{2} V_{jj'}^{\alpha}(\boldsymbol{q}) V_{j'j}^{\beta}(\boldsymbol{q}) \frac{ ( \frac{\partial n_{\boldsymbol{q}j}}{\partial T} + \frac{\partial n_{\boldsymbol{q}j'}}{\partial T})(\Gamma_{\boldsymbol{q}j}+\Gamma_{\boldsymbol{q}j'}) }{4(\omega_{\boldsymbol{q}j}-\omega_{\boldsymbol{q}j'})^2 + (\Gamma_{\boldsymbol{q}j}+\Gamma_{\boldsymbol{q}j'})^2}\]

where \(k^{BTE}_{\alpha\beta}\) is the thermal conductivity estimated by the Boltzmann transport equation discussed above, and \(\Gamma_{\boldsymbol{q}j} = \frac{1}{\tau_{\boldsymbol{q}j}}\) is the phonon linewidth, i.e. a diagonal element of the scattering matrix.

Thermal Viscosity

The theory of the phonon thermal viscosity is described to far greater extent in this reference https://journals.aps.org/prx/abstract/10.1103/PhysRevX.10.011019.

In short, the equilibrium distribution for a system of bosonic particles that conserves energy and momentum is the drifting distribution,

\[n_{\nu}^{D} = \frac{1}{e^{\beta(\hbar \omega_\nu - \hbar \boldsymbol{q} \cdot \boldsymbol{u})}-1} \;,\]

where \(\boldsymbol{q}\) is the phonon wavevector (proportional to the phonon crystal momentum, and \(\boldsymbol{u}\) is the phonon drift velocity. The thermal viscosity is defined as the coefficient of proportionality between the crystal momentum flux \(\Pi\) and a local perturbation in the drift velocity \(\boldsymbol{u}\),

\[\Pi^{ij} = - \sum_{kl} \eta^{ijkl} \frac{\partial u^k}{\partial r^l}\]

and the momentum flux (at least, the component relevant to our case) is defined as:

\[\Pi^{ij} = \frac{1}{V N_q} \sum_{\nu} \hbar q^i v_{\nu}^j n_{\nu}\]

The population in response to the perturbation is fixed by the phonon BTE. At the RTA level, we simply need to solve

\[\boldsymbol{v}_{\nu} \cdot \left(\frac{\partial n^{D}_{\nu}}{\partial \boldsymbol{u}} \cdot \nabla \boldsymbol{u} \right) = - \frac{n_{\nu}}{\tau_{\nu}}\]

We linearize the solution, stating \(n_{\nu} = n_{\nu} \nabla \boldsymbol{u}\), and the equation is readily solved.

Beyond the RTA (the relaxons solver case), we must solve the equation,

\[\frac{\boldsymbol{v}_{\nu}}{\sqrt{\bar{n}_{\nu}(\bar{n}_{\nu}+1)}} \cdot \left(\frac{\partial n^{D}_{\nu}}{\partial \boldsymbol{u}} \cdot \nabla \boldsymbol{u} \right) = - \frac{1}{V N_q} \sum_{\nu'} \Omega_{\nu\nu'} n_{\nu'}\]

which we do with the eigenvector formalism. Using the eigenvectors of the scattering matrix, we expand the phonon population as:

\[n_{\nu} = \sum_{kl} f^{kl}_{\alpha} \theta_{\nu\alpha} \nabla_l u^k\]

We find the solution as:

\[f^{kl}_{\alpha} = - \tau_{\alpha} \sum_{\nu} \theta_{\nu\alpha} \frac{\boldsymbol{v}^l_{\nu}}{\sqrt{\bar{n}_{\nu}(\bar{n}_{\nu}+1)}} \frac{\partial n^{D}_{\nu}}{\partial u^k}\]

which can be used to reconstruct the phonon population response. Finally, the viscosity tensor is symmetrized, finding the thermal viscosity:

\[\mu^{ijkl} = \frac{1}{2} \left( \eta^{ijkl} + \eta^{ilkj} \right)\]

The code also prints other quantities needed to write the viscous heat equations derived in this reference https://journals.aps.org/prx/abstract/10.1103/PhysRevX.10.011019.

Symmetries of the BTE

We exploit the symmetries of the crystal to speed up the calculation of thermal conductivity. We primarily follow the reference https://journals.aps.org/prl/abstract/10.1103/PhysRevLett.110.265506.

Let \(q\) indicate any wavevector in the Brillouin zone. The symmetries of a crystal identify an irreducible set of wavevectors \(q^*\), such that any other wavevector \(q\) can be obtained from a rotation of these irreducible wavevectors \(q = R q^*\). The basic idea is to restrict the calculation to the irreducible set of wavevectors. The conductivity for example, is:

\[k^{ij} = \frac{1}{V N_k} \sum_{\nu} \hbar \omega_{\nu} v^i_{\nu} n^{j}_{\nu} = \frac{1}{V N_k} \sum_{\nu^*} \sum_{R} \hbar \omega_{\nu^*} (R v_{\nu})_{i} (R n_{\nu})_{j}\]

where \(R\) is the set of rotations used to reconstruct all the symmetry-equivalent wavevectors of \(q^*\), and the summation over \(\nu^*\) is only done in the irreducible set of wavevectors.

The BTE too can be restricted to the irreducible wedge.

\[v^i_{\nu^*} \frac{\partial \bar{n}_{\nu}}{\partial T} = - \frac{1}{V N_q} \sum_{\nu'} A_{\nu^*\nu'} f^i_{\nu'} = - \frac{1}{V N_q} \sum_{\nu'^*} \sum_{R} \sum_{j} A_{\nu^*\nu'^*} R_{ij} f^j_{\nu'} = - \frac{1}{V N_q} \sum_{\nu'^* j} A^{ij}_{\nu^*\nu'^*} f^j_{\nu'}\]

Hence, one can work with the same techniques detailed above, provided that we work with an enlarged matrix \(A^{ij}_{\nu^*\nu'^*}\).

Some comments:

  • Advantage: for a system with a lot of symmetries, the matrix \(A^{ij}_{\nu^*\nu'^*}\) is generally smaller than \(A_{\nu\nu'}\), and thus calculations will be much faster.

  • Disadvantage 1: we cannot compute viscosity beyond the RTA using symmetries. To do so, one must disable symmetries.

  • Disadvantage 2: note that the symmetric matrix gains two Cartesian coordinate indices. As a result, in the limiting case of no symmetries in the system (only the identity), the matrix \(A^{ij}_{\nu^*\nu'^*}\) will still be computed on the same number of wavevectors of \(A_{\nu\nu'}\), but occupies 3x3 times more memory without adding any information. Therefore, for low-symmetry systems, consider disabling symmetries.

  • Disadvantage 3: The symmetries of the BTE are so far not applicable to the variational and relaxons solveres. This is not so much a problem with implementaiton, but instead is because of a need for a derivation of symmetries for these cases.