next up previous contents
Next: Density Functional Theory Up: Techniques Previous: Reaction rate theory   Contents


Car-Parrinello molecular dynamics

The motion of small, but not too small, particles such as atoms and molecules is usually well described by Lagrangian mechanics, with the Lagrangian ${\cal L}$ defined as the kinetic energy minus the potential energy,

\begin{displaymath}
{\cal L} = T-V
\end{displaymath} (21)

leading to the set of Newtonian equations of motion for each particle $\alpha $, with mass $m_\alpha$ and cartesian coordinate ${\bf R}_\alpha $:
\begin{displaymath}
m_\alpha \frac{\mathrm{d}^2 {\bf R}_\alpha }{\mathrm{d}t^2} =
- \frac{\partial V}{\partial {\bf R}_\alpha }
\end{displaymath} (22)

Numerical integration of the equations of motion, taking small time steps $\Delta t$, results in a trajectory through the hyper-space of all possible positions and momenta of the particles, called phase-space. This simulation technique is known as molecular dynamics (MD).

The motion of even smaller particles, such as electrons, cannot be approximated with classical Newtonian dynamics, but instead has to be described with the more accurate quantum mechanical equations of motion derived from the time-independent Schrödinger equation:

\begin{displaymath}
{\hat {\cal H}} \Psi = E \Psi
\end{displaymath} (23)

That is, the Hamiltonian ${\hat {\cal H}}$, given as the sum of the kinetic and potential energy operators, operating on the many-electron wave function $\Psi$ gives us the energy $E$. This rather simple and very famous equation can unfortunately not be solved analytically for a system of more than two electrons, and approximations on the Hamiltonian have to be introduced. Following Kohn and Sham's density functional theory (DFT) approach (see next section for the details on DFT), the exact density and the exact energy can be obtained from one-electron wave functions, the Kohn-Sham (KS) orbitals $\psi_i$, which are solutions in a local potential. The heart of this machinery lies in the notion that all properties, including the energy, of the electronic system are a functional of the electron density, $\rho$,
\begin{displaymath}
E=E[\rho]
\end{displaymath} (24)

The Kohn-Sham orbitals can be expanded in an orthogonal basis $\chi_1, \chi_2, \ldots$:
\begin{displaymath}
\psi_i = \sum_k c_k^i \chi_k
\end{displaymath} (25)

with $c_k^i$ the expansion coefficients. The electron density $\rho$, given as
\begin{displaymath}
\rho({\bf r}) = \sum_i f_i \left\vert \psi_i({\bf r}) \right\vert^2
\end{displaymath} (26)

with $f_i$ the occupation number of KS orbital $i$, gives us the probability to find an electron at position ${\bf r}$.

In DFT, the sets of coefficients $c_k^i$ span a hyper-space $\cal C$, in which a point representing orthonormal KS orbitals $\psi_i$ thus corresponds to an energy $E$, similar to a point in atomic coordinate space corresponding to a potential energy $V({\bf R}^N)$. In fact, the iterative search for the one point in $\cal C$ that minimizes the energy to the physical ground-state energy $E_0$, resembles a molecular geometry optimization to minimize the potential energy. Since we want the one-electron functions $\psi_i$ to be orthonormal, the trajectory through $\cal C$ during the energy minimization is constrained on a hyper-surface in $\cal C$ for which

\begin{displaymath}
\sum_k c^{i*}_kc^j_k=\delta_{ij} \quad \mbox{with}
\quad \de...
...x{for $i=j$} \\
0 & \mbox{for $i \neq j$}
\end{array} \right.
\end{displaymath} (27)

Taking the iteration steps as the analog of time steps, we can even formulate Newtonian equations of motion for the ``dynamics'' of changing coefficients:
\begin{displaymath}
\mu_k \frac{\mathrm{d}^2 c^i_k}{\mathrm{d}t^2} =
- \frac{\partial E}{\partial c^{i*}_k} - \sum_j \lambda_{ij} c^j_k
\end{displaymath} (28)

Here, the $\mu_k$ are inertia parameters, usually called ``fictitious masses'', which control the acceleration of the coefficients, $- \frac{\mathrm{d}^2 c^i_k}{\mathrm{d}t^2}$ due to the force on the coefficients, on the right-hand-side. The last term arises from the constraint in equation 2.26, with $\lambda_{ij}$ the undetermined Lagrange multipliers. The forces on the coefficients are given as
\begin{displaymath}
- \frac{\partial E[\{c_k^i\},\{ {\bf R}_\alpha \}] }{\partial c_k^i } =
- \sum_l n_i H^\mathrm{KS}_{kl} c_l^i
\end{displaymath} (29)

which can amount to a considerable saving of computer time and memory with respect to techniques based on diagonalization of the full Hamiltonian matrix. In practice, the wave function coefficients are not known a priori, and we start from a random set of $c_k^i$, consistent with equation 2.26, associated with a meaningless energy $E$ via equations 2.23 and 2.24. When we start to integrate the equations of motion (eq. 2.27), the coefficients will accelerate towards configurations with lower energies and gain ``kinetic energy'', until a dynamic equilibrium is reached. The electronic ground-state energy $E_0$ is found by damping the coefficient dynamics, so that kinetic energy is gradually removed and the coefficients eventually freeze in the ground-state configuration. This technique to optimize the wave function is known as ``simulated annealing''.

Car-Parrinello molecular dynamics (CPMD) is the integration of the fictitious wave function coefficient dynamics with the classical molecular dynamics by a single extended Lagrangian.

$\displaystyle {\cal L}$ $\textstyle =$ $\displaystyle \sum_\alpha \frac{1}{2} m_\alpha
\left\vert \frac{\mathrm{d}{\bf ...
... \frac{1}{2} \mu_k
\left\vert \frac{\mathrm{d}c_k^i}{\mathrm{d}t} \right\vert^2$  
  $\textstyle -$ $\displaystyle E[\{c_k^i\},\{{\bf R}_\alpha \}]
+ \sum_{i,j} \lambda_{ij} (\sum_{k,l} c^i_l c^j_k S_{kl} - \delta_{ij})$ (30)

Starting from some atomic configuration in the (optimized) electronic ground-state, we can calculate the forces on the atoms using the Hellmann-Feynman theorem,
\begin{displaymath}
- \frac{\partial E[\{c_k^i\},\{ {\bf R}_\alpha \}] }{\partia...
... {\cal H}}{\partial {\bf R}_\alpha }
\left\vert \Psi \right >
\end{displaymath} (31)

to start the ab initio molecular dynamics. Initially, the forces on the coefficients equal zero as the electronic configuration is at its minimum, but after one time step the atomic positions have changed and the wave function is no longer up to date. However, since the electronic degrees of freedom $c_k^i$ and the atomic positions ${\bf R}_\alpha $ are coupled via the potential energy (equation 2.29), the forces on the coefficients are no longer zero and the coefficients accelerate towards the new electronic ground-state. When the fictitious coefficient masses $\mu_k$ are chosen sufficiently small (i.e. $\mu_k\ll m_\alpha \forall k, \alpha$) the response of the coefficients to the changing nuclei $\alpha $ is so rapid that the electrons remain to a sufficiently high degree in the ground-state.

Figure 2.2: Vibrational spectrum of the normal modes of the electronic coefficients (solid line) for a system with a large gap (periodic super cell containing 8 Si atoms in the diamond structure). The solid triangle indicates the position of the highest ionic frequency. The vertical bars below the spectrum represent the frequencies obtained from equation 2.31. Figure from ref PaSmBu91.

Of course, the dynamics of the nuclei in CPMD has only physical meaning if the electronic structure is close to its instantaneous ground-state at each step of the simulation. In other words, the dynamics of the electronic coefficients $c_k^i$ has to remain relatively cold. However, since the two dynamic sub-systems of nuclei ${\bf R}_\alpha $ and electronic coefficients $c_k^i$ are coupled, in principle energy can flow from the relatively hot nuclei sub-system, to the colder coefficients, which would lead to deviations from the Born-Oppenheimer (ground-state) surface. In many practical simulations, the energy flow between the sub-systems can be suppressed with a good choice for the fictitious coefficient masses $\mu_k$, which can be rationalized by regarding the electronic coefficient dynamics for small deviations from the ground-state, described as a superposition of harmonic oscillations whose frequency is given by:

$\displaystyle \omega^{(1)}_{ij}$ $\textstyle =$ $\displaystyle [f_j(\epsilon_i - \epsilon_j)/\mu]^{1/2}$  
$\displaystyle \omega^{(2)}_{ij}$ $\textstyle =$ $\displaystyle [(f_j-f_i)(\epsilon_i - \epsilon_j)/2\mu]^{1/2}$ (32)

Here, $\epsilon_i$ indicates the eigenvalue of the $i$th unoccupied and $\epsilon_j$ the $j$th occupied level, and the fictitious coefficient masses $\mu_k$ are chosen equal for all $k$. As an illustration, let us refer to figure 2.2 from the illuminating study of ref PaSmBu91, which shows the vibrational density of states of the electronic coefficients for an unrealistic but instructive model of crystalline silicon. The fictitious mass in this work was $\mu=300$ au. The solid line is obtained from the Fourier transform of the velocity autocorrelation function
\begin{displaymath}
\nu(\omega) = \int^\infty_0 \mathrm{d}t \cos{(\omega t)}
\sum_{k,i} \left< \dot{c}_k^i(t) \dot{c}_k^i(0) \right>
\end{displaymath} (33)

from a simulation of 3000 time steps ($\Delta t=5$ au) and the vertical bars below are obtained from equation 2.31. The lowest electronic frequency (at about 1010 THz) results from the energy gap, which is in this model $E_g=2.24$ eV, in good agreement with the estimate using equation 2.31: $\omega_\mathrm{min}=({2E_g/\mu})^{1/2}=968$ THz. The highest ionic frequency for this model is 140 THz, indicated by the triangle in figure 2.2. This clear separation between the characteristic electronic and ionic frequencies, is the reason that the irreversible energy transfer from the slow to the fast degrees of freedom is minimal, and the main reason that CPMD works! Problems occur for systems with a small or vanishing bandgap, such as semiconductors and metals, because the heat transfer can no longer be controlled by choosing a small enough $\mu$. A workable solution for these systems can be provided by coupling both dynamical sub-systems to thermostats which remove kinetic energy from the electronic coefficients while adding kinetic energy to the nuclei[19].

As the electronic coefficients rapidly fluctuate around their optimal values, the instantaneous values of the forces do not coincide with the Hellmann-Feynmann forces, however their average values do to a very high degree of accuracy. The high efficiency of the Car-Parrinello approach with respect to the so-called (real-) Born-Oppenheimer molecular dynamics (BOMD) method also lies in this respect. In BOMD, the electronic structure is self-consistently optimized after each time step in which only the relatively small number of nuclei positions are propagated. In practice, the convergence of the optimization has to be very high to avoid the accumulation of the small, but systematic (!), deviations in the Hellmann-Feynman forces, which makes BOMD computational more demanding than CPMD. This difference in efficiency is partly reduced by the larger time step that can be used in BOMD, as the maximum time step is limited by the proper integration of the equation of motions of the fastest (i.e. lightest) particles, which are the high frequency components of the fictitious dynamics in CPMD, whereas in BOMD, the lightest nuclei limit the time step maximum. A technique to accelerate CPMD further is known as mass-preconditioning[20], which is based on reducing these high frequency components of the fictitious dynamics by properly scaling the fictitious masses $\mu_k$, so that a larger time step can be used. Excellent reviews of the Car-Parrinello molecular dynamics techniques are found in refs MaHuReview,ReMaReview,GaPaReview.

Figure 2.3: Partial waves and projectors for Mn. Left panel: partial waves $\phi $ (solid lines) and pseudo partial waves $\tilde{\phi}$ (dashed and dash-dotted lines). The ``first'' pseudo partial wave is a dash-dotted line. Right panel: first (solid line) and second (dashed line) projector functions. (a) and (d) show the results for the first and second partial wave of the $s$ angular momentum channel, respectively, (b) and (e) for the $p$ channel and (c) and (f) for the $d$ channel. 3$s$ and 3$p$ functions are treated as valence states. Figure from ref PAW.

The first CPMD implementations used a basis set of plane waves, in combination with a pseudopotential[25,26,27], to expand the one-electron valence wave functions, which is the most widely used approach in electronic structure calculations in the field of solid state physics. Plane waves are of the form $\psi=N^{-1}\exp{[i{\bf Gr}]}$ ($i=(-1)^{1/2}$ and $G$ the wave vector), which are the eigenfunctions of Schrödinger's equation (2.22) for an electron in vacuum with kinetic energy $E_\mathrm{kin}=G^2/2$. For electrons in an external field, such as in atoms, however, an expansion in plane waves is a rather poor choice, because these functions hardly mimic the rapid fluctuations of the one-electron wave functions in the neighborhood of the nucleus. Usually the low lying core electrons are kept "frozen" during an electronic structure calculation to reduce the computational cost, which is a good approximation since these states remain practically invariant from their atomic ground-state during the chemistry of bond breaking and making. For the chemically active valence wave functions a pseudopotential is used to describe the inner region (nucleus plus core electrons) of the atom, in a way that allows for replacing the rapid fluctuating functions by more smooth functions. Beyond this inner region, the smooth functions agree with the true wave functions. A disadvantage of the pseudopotential method is that they become "very hard" for first row elements and for systems with $d$ and $f$ electrons, so that very large basis sets are required. An improvement on the traditional pseudopotentials was provided with the development of Vanderbilt's ultrasoft pseudopotentials [28,29] by relaxing on the norm conserving condition that is usually imposed. The simulations described in this thesis were performed with Blöchl's CPMD implementation, named projector augmented wave or PAW[24], which is based on a generalization of both the pseudopotential approach and the linear augmented-plane wave method[30,31,32] (LAPW). In contrast to the pseudopotential approach, the LAPW technique utilizes the full one-electron valence wave function with the correct nodal structure. The wave function is subdivided into two regions; an inner (augmentation) region, which is expanded in a basis set of localized functions, and an outer region which is described by plane waves. At a certain (muffin-tin) sphere radius from the nucleus, the partial wave functions are matched together by value and derivative of the functions. Instead in PAW, a linear transformation between the one-electron valence wave functions $\psi$ and fictitious pseudo wave functions $\tilde{\psi}$ is used, in Dirac's bra and ket notation:

\begin{displaymath}
\vert\psi> = \tilde{\psi} + \sum_i (\vert\phi_i> - \vert\tilde{\phi}_i>)<\tilde{p}_i\vert\tilde{\psi}>
\end{displaymath} (34)

where $\vert\phi>_i$ are a complete set of partial waves ($i$ referring to the atomic site $R$, the angular momentum quantum numbers $L=(l,m)$ and the index $n$ for different partial waves per $R$ and $L$). Each partial wave $\vert\phi>_i$ is connected to a pseudo partial wave $\vert\tilde{\phi}>_i$, which only differs from $\vert\phi>_i$ inside an augmentation region $\Omega_R$, and a localized, so-called, projector function $<\tilde{p}_i\vert$ for which,
\begin{displaymath}
<\tilde{p}_i\vert\tilde{\phi}> = \delta_{ij}.
\end{displaymath} (35)

As an illustration, we use figure 2.3 from ref PAW, which shows the partial waves and projector functions for manganese. The partial waves are functions on a radial grid, multiplied with spherical harmonics, which are obtained by radially integrating Schrödinger's equation for the isolated atom. The pseudo wave functions $\tilde{\psi}$ are expanded in plane waves. In practice, the partial waves $\vert\phi_i>$ and $\vert\tilde{\phi}_i>$ and the projector functions $<\tilde{p}_i\vert$ are imported from an atomic calculation at the start of a PAW molecular dynamics calculation. The main advantages of the PAW approach compared to the conventional norm-conserving pseudopotential methods is that the atomic partial and projector functions are better transferable than pseudopotentials and that the full wave function is accessible, which allows for the calculation of core dependent quantities such as hyperfine parameters and electric field gradients[33,34]. Also a smaller plane wave basis set is required than with the norm conserving pseudopotential methods, which is most important for the present study as it cuts heavily on the computational expense. A very detailed description of the PAW technique and its relation to the traditional norm conserving pseudopotential method and the LAPW method is found in reference PAW.


next up previous contents
Next: Density Functional Theory Up: Techniques Previous: Reaction rate theory   Contents
Bernd Ensing 2003-06-13