X-Git-Url: https://hackdaworld.org/gitweb/?p=lectures%2Flatex.git;a=blobdiff_plain;f=physics_compact%2Fsolid.tex;h=1fc7a17eea47e2ac231e10db76dfc7321a9e3c2c;hp=50df24d29d03741e8a4efb3933ae5b98316dcbba;hb=f47fcf0c7b7c7e2d5adc294d72a0c914289c584c;hpb=9fac15a186e73cf587fb7d9dfa1c737c778cd08a diff --git a/physics_compact/solid.tex b/physics_compact/solid.tex index 50df24d..1fc7a17 100644 --- a/physics_compact/solid.tex +++ b/physics_compact/solid.tex @@ -2,6 +2,10 @@ \chapter{Atomic structure} +\chapter{Reciprocal lattice} + +Example of primitive cell ... + \chapter{Electronic structure} \section{Noninteracting electrons} @@ -18,11 +22,58 @@ \subsubsection{Hohenberg-Kohn theorem} -Considering a system with a nondegenerate ground state, there is obviously only one ground-state charge density $n_0(\vec{r})$ that correpsonds to a given potential $V(\vec{r})$. +The Hamiltonian of a many-electron problem has the form +\begin{equation} +H=T+V+U\text{ ,} +\end{equation} +where +\begin{eqnarray} +T & = & \langle\Psi|\sum_{i=1}^N\frac{-\hbar^2}{2m}\nabla_i^2|\Psi\rangle\\ + & = & \frac{-\hbar^2}{2m} \sum_{i=1}^N \int d\vec{r} d\vec{r}' \, + \langle \Psi | \vec{r} \rangle \langle \vec{r} | + \nabla_i^2 + | \vec{r}' \rangle \langle \vec{r}' | \Psi \rangle\\ + & = & \frac{-\hbar^2}{2m} \sum_{i=1}^N \int d\vec{r} d\vec{r}' \, + \langle \Psi | \vec{r} \rangle \nabla_{\vec{r}_i} + \langle \vec{r} | \vec{r}' \rangle + \nabla_{\vec{r}'_i} \langle \vec{r}' | \Psi \rangle\\ + & = & \frac{-\hbar^2}{2m} \sum_{i=1}^N \int d\vec{r} d\vec{r}' \, + \nabla_{\vec{r}_i} \langle \Psi | \vec{r} \rangle + \delta_{\vec{r}\vec{r}'} + \nabla_{\vec{r}'_i} \langle \vec{r}' | \Psi \rangle\\ + & = & \frac{-\hbar^2}{2m} \sum_{i=1}^N \int d\vec{r} \, + \nabla_{\vec{r}_i} \Psi^*(\vec{r}) \nabla_{\vec{r}_i} \Psi(\vec{r}) + \text{ ,} \\ +V & = & \int V(\vec{r})\Psi^*(\vec{r})\Psi(\vec{r})d\vec{r} \text{ ,} \\ +U & = & \frac{1}{2}\int\frac{1}{\left|\vec{r}-\vec{r}'\right|} + \Psi^*(\vec{r})\Psi^*(\vec{r}')\Psi(\vec{r}')\Psi(\vec{r}) + d\vec{r}d\vec{r}' +\end{eqnarray} +represent the kinetic energy, the energy due to the external potential and the energy due to the mutual Coulomb repulsion. + +\begin{remark} +As can be seen from the above, two many-electron systems can only differ in the external potential and the number of electrons. +The number of electrons is determined by the electron density. +\begin{equation} +N=\int n(\vec{r})d\vec{r} +\end{equation} +Now, if the external potential is additionally determined by the electron density, the density completely determines the many-body problem. +\end{remark} + +Considering a system with a nondegenerate ground state, there is obviously only one ground-state charge density $n_0(\vec{r})$ that corresponds to a given potential $V(\vec{r})$. +\begin{equation} +n_0(\vec{r})=\int \Psi_0^*(\vec{r},\vec{r}_2,\vec{r}_3,\ldots,\vec{r}_N) + \Psi_0(\vec{r},\vec{r}_2,\vec{r}_3,\ldots,\vec{r}_N) + d\vec{r}_2d\vec{r}_3\ldots d\vec{r}_N +\end{equation} In 1964, Hohenberg and Kohn showed the opposite and far less obvious result \cite{hohenberg64}. + +\begin{theorem}[Hohenberg / Kohn] For a nondegenerate ground state, the ground-state charge density uniquely determines the external potential in which the electrons reside. -The proof presented by Hohenberg and Kohn proceeds by {\em reductio ad absurdum}. +\end{theorem} +\begin{proof} +The proof presented by Hohenberg and Kohn proceeds by {\em reductio ad absurdum}. Suppose two potentials $V_1$ and $V_2$ exist, which yield the same electron density $n(\vec{r})$. The corresponding Hamiltonians are denoted $H_1$ and $H_2$ with the respective ground-state wavefunctions $\Psi_1$ and $\Psi_2$ and eigenvalues $E_1$ and $E_2$. Then, due to the variational principle (see \ref{sec:var_meth}), one can write @@ -55,5 +106,169 @@ E_1 + E_2 < E_2 + E_1 + \int n(\vec{r}) \left( V_2(\vec{r})-V_1(\vec{r}) \right) d\vec{r} }_{=0} \end{equation} -is revealed. +is revealed, which proofs the Hohenberg Kohn theorem.% \qed +\end{proof} + +\section{Electron-ion interaction} + +\subsection{Pseudopotential theory} + +The basic idea of pseudopotential theory is to only describe the valence electrons, which are responsible for the chemical bonding as well as the electronic properties for the most part. + +\subsubsection{Orthogonalized planewave method} + +Due to the orthogonality of the core and valence wavefunctions, the latter exhibit strong oscillations within the core region of the atom. +This requires a large amount of planewaves $\ket{k}$ to adequatley model the valence electrons. + +In a very general approach, the orthogonalized planewave (OPW) method introduces a new basis set +\begin{equation} +\ket{k}_{\text{OPW}} = \ket{k} - \sum_t \ket{t}\bra{t}k\rangle \text{ ,} +\end{equation} +with $\ket{t}$ being the eigenstates of the core electrons. +The new basis is orthogonal to the core states $\ket{t}$. +\begin{equation} +\braket{t}{k}_{\text{OPW}} = +\braket{t}{k} - \sum_{t'} \braket{t}{t'}\braket{t'}{k} = +\braket{t}{k} - \braket{t}{k}=0 +\end{equation} +The number of planewaves required for reasonably converged electronic structure calculations is tremendously reduced by utilizing the OPW basis set. + +\subsubsection{Pseudopotential method} + +Following the idea of orthogonalized planewaves leads to the pseudopotential idea, which --- in describing only the valence electrons --- effectively removes an undesriable subspace from the investigated problem. + +Let $\ket{\Psi_\text{V}}$ be the wavefunction of a valence electron with the Schr\"odinger equation +\begin{equation} +H \ket{\Psi_\text{V}} = \left(\frac{1}{2m}p^2+V\right)\ket{\Psi_\text{V}}= +E\ket{\Psi_\text{V}} \text{ .} +\end{equation} +\ldots projection operatore $P_\text{C}$ \ldots + +\subsubsection{Semilocal form of the pseudopotential} + +Ionic potentials, which are spherically symmteric, suggest to treat each angular momentum $l,m$ separately leading to $l$-dependent non-local (NL) model potentials $V_l(r)$ and a total potential +\begin{equation} +V=\sum_{l,m}\ket{lm}V_l(r)\bra{lm} \text{ .} +\end{equation} +In fact, applied to a function, the potential turns out to be non-local in the angular coordinates but local in the radial variable, which suggests to call it asemilocal (SL) potential. + +Problem of semilocal potantials become valid once matrix elements need to be computed. +Integral with respect to the radial component needs to be evaluated for each planewave combination, i.e.\ $N(N-1)/2$ integrals. +\begin{equation} +\bra{k+G}V\ket{k+G'} = \ldots +\end{equation} + +A local potential can always be separated from the potential \ldots +\begin{equation} +V=\ldots=V_{\text{local}}(r)+\ldots +\end{equation} +\subsubsection{Norm conserving pseudopotentials} + +HSC potential \ldots + +\subsubsection{Fully separable form of the pseudopotential} + +KB transformation \ldots + +\subsection{Spin-orbit interaction} + +Relativistic effects can be incorporated in the normconserving pseudopotential method up to but not including order $\alpha^2$ with $\alpha$ being the fine structure constant. +This is advantageous since \ldots +With the solutions of the all-electron Dirac equations, the new pseudopotential reads +\begin{equation} +V(r)=\sum_{l,m}\left[ +\ket{l+\frac{1}{2},m+{\frac{1}{2}}}V_{l,l+\frac{1}{2}}(r) +\bra{l+\frac{1}{2},m+{\frac{1}{2}}} + +\ket{l-\frac{1}{2},m-{\frac{1}{2}}}V_{l,l-\frac{1}{2}}(r) +\bra{l-\frac{1}{2},m-{\frac{1}{2}}} +\right] \text{ .} +\end{equation} +By defining an averaged potential weighted by the different $j$ degeneracies of the $\ket{l\pm\frac{1}{2}}$ states +\begin{equation} +\bar{V}_l(r)=\frac{1}{2l+1}\left( +l V_{l,l-\frac{1}{2}}(r)+(l+1)V_{l,l+\frac{1}{2}}(r)\right) +\end{equation} +and a potential describing the difference in the potential with respect to the spin +\begin{equation} +V^{\text{SO}}_l(r)=\frac{2}{2l+1}\left( +V_{l,l+\frac{1}{2}}(r)-V_{l,l-\frac{1}{2}}(r)\right) +\end{equation} +the total potential can be expressed as +\begin{equation} +V(r)=\sum_l \ket{l}\left[\bar{V}_l(r)+V^{\text{SO}}_l(r)LS\right]\bra{l} +\text{ ,} +\end{equation} +where the first term correpsonds to the mass velocity and Darwin relativistic corrections and the latter is associated with the spin-orbit (SO) coupling. + + +\subsubsection{Excursus: real space representation suitable for an iterative treatment} + +In the following, the spin-orbit part is evaluated in real space. +Since spin is treated in another subspace, it can be treated separately. +The matrix elements of the orbital angular momentum part of the potential in KB form for orbital angular momentum $l$ read +\begin{equation} +\bra{r'}(r\times p)\ket{\chi_{lm}}E^{\text{SO,KB}}_l\braket{\chi_{lm}}{r''} +\text{ .} +\end{equation} +With +\begin{eqnarray} +\bra{r'}p\ket{\chi_{lm}} & = & -i\hbar\nabla_{r'} \braket{r'}{\chi_{lm}} +=-i\hbar\nabla_{r'}\,\chi_{lm}(r') \\ +r\ket{r'} & = & r'\ket{r'} +\end{eqnarray} +we get +\begin{equation} +-i\hbar(r'\times \nabla_{r'})\braket{r'}{\chi_{lm}}E^{\text{SO,KB}}_l +\braket{\chi_{lm}}{r''} +\text{ .} +\end{equation} +To further evaluate this expression, the KB projectors +\begin{equation} +\chi_{lm}=\frac{\ket{\delta V_l^{\text{SO}}\Phi_{lm}}} +{\braket{\delta V_l^{\text{SO}}\Phi_{lm}} + {\Phi_{lm}\delta V_l^{\text{SO}}}^{1/2}} +\end{equation} +must be known in real space (with respect to $r$). +\begin{equation} +\braket{r'}{\chi_{lm}}= +\frac{\braket{r'}{\delta V_l^{\text{SO}}\Phi_{lm}}}{ +\braket{\delta V_l^{\text{SO}}\Phi_{lm}}{\Phi_{lm}\delta V_l^{\text{SO}}} +^{1/2}} +\end{equation} +and +\begin{equation} +\braket{r'}{\delta V_l^{\text{SO}}\Phi_{lm}}= +\delta V_l^{\text{SO}}(r')\frac{u_l(r')}{r'}Y_{lm}(\Omega_{r'}) +\text{ .} +\end{equation} +In this expression, only the spherical harmonics are complex functions. +Thus, the complex conjugate with respect to $r''$ is given by +\begin{equation} +\braket{\Phi_{lm}\delta V_l^{\text{SO}}}{r''}= +\delta V_l^{\text{SO}}(r'')\frac{u_l(r'')}{r''}Y^*_{lm}(\Omega_{r''}) +\text{ .} +\end{equation} +Using the orthonormality property +\begin{equation} +\int Y^*_{l'm'}(\Omega_r)Y_{lm}(\Omega_r) d\Omega_r = \delta_{ll'}\delta_{mm'} +\label{eq:solid:y_ortho} +\end{equation} +of the spherical harmonics, the norm of the $\chi_{lm}$ reduces to +\begin{eqnarray} +\braket{\delta V_l^{\text{SO}}\Phi_{lm}}{\Phi_{lm}\delta V_l^{\text{SO}}} &=& +\int \braket{\delta V_l^{\text{SO}}\Phi_{lm}}{\vec{r}'} +\braket{\vec{r}'}{\Phi_{lm}\delta V_l^{\text{SO}}} d\vec{r'}'\\ +&=&\int +{\delta V_l^{\text{SO}}}^2(r')\frac{u_l^2(r')}{r'^2}Y^*_{lm}(\Omega_{r'}) +Y_{lm}(\Omega_{r'}) +r'^2 dr' d\Omega_{r'} \\ +&=&\int_{r'} +{\delta V_l^{\text{SO}}}^2(r') u_l^2(r') dr' +\int_{\Omega_{r'}}Y^*_{lm}(\Omega_{r'})Y_{lm}(\Omega_{r'}) d\Omega_{r'}\\ +&=&\int_{r'}{\delta V_l^{\text{SO}}}^2(r') u_l^2(r') dr' \text{ .} +\end{eqnarray} + +Finally, to evaluate $V^{\text{SO}}_l\ket{\Psi}$, the integral \ldots +\begin{equation} +\end{equation}