next up previous
Next: Conclusions Up: Nonorthogonal generalized Wannier function Previous: The FFT box technique


The NGWF pseudopotential plane-wave method in practice

We have implemented our method in a new code and we have performed extensive tests on a variety of systems. We have also performed comparisons with CASTEP [1], an established pseudopotential plane-wave code that we use as our point of reference. We expect our approach to have efficiency comparable to the traditional cubic-scaling plane-wave pseudopotential method and thus it would be possible to use it in its place for systems with a band gap. In that case, a calculation with our method would afford a set of optimal localised functions which could be used directly in applications such as the calculation of polarisation changes in crystalline solids [48,49]. However, the most important application that we envisage is the extension of the present formalism to linear-scaling calculations on very large systems. Such an extension requires the truncation of the density kernel, an issue which has been already investigated in detail [33,35,50]. The resulting linear-scaling method would be directly comparable to and have the same advantages as the plane-wave approach.

Since we optimise the NGWFs iteratively, some initial guesses are required for them. In this work we use pseudo-atomic orbitals (PAOs) that vanish outside a spherical region [51]. These orbitals are generated for the isolated atoms with the same radii, norm-conserving pseudopotentials and kinetic energy cutoff as in our calculations. Even though these NGWF guesses are optimal for the isolated atoms, they undergo large changes during our calculations so that in practice any guess that resembles an atomic orbital could be used, such as Slater or Gaussian functions.

We first demonstrate the accuracy of the FFT box technique as compared to using the entire simulation cell as the FFT grid. We define the quantity

\begin{displaymath}
\Delta E \equiv E^{\mathrm{box}}[n] - E[n],
\end{displaymath} (30)

where $E^{\mathrm{box}}$ is the total energy per atom calculated using the FFT box technique and $E$ is that calculated using the entire simulation cell. Figure 2 shows $\Delta E$ for the butane molecule (C$_{4}$H$_{10}$) for different FFT box sizes. For this test we used a cubic simulation cell of side length $50a_0$ and grid spacing $0.5a_0$. The PAOs on all the atoms were confined within spherical regions of radius $6.0a_0$. The carbon atoms had one 2s and three 2p orbitals and the hydrogen atoms had a single 1s orbital. In this case the PAOs were not optimised during the calculation.

Figure 2: $\Delta E$ plotted for a butane molecule as a function of FFT box size. All PAOs were confined to atom centered localisation regions of radius $6.0a_0$, and the grid spacing was $0.5a_0$.

\scalebox{0.5}{\includegraphics*[0cm,1cm][26cm,18.5cm]{butane.eps} }

It is seen that the error associated with using the FFT box rather than the entire simulation cell is only of the order of $10^{-7}$E$_h$ per atom, which is insignificant in the context of DFT calculations. We also note that the convergence of the total energy with FFT box size is not strictly variational, as is expected: as the FFT box size is increased, it is true that the basis set expands, but the smaller basis is not necessarily a subset of the larger one. For a given FFT box size, however, the kinetic energy cutoff of our basis functions (and hence the grid-spacing) is a variational parameter, just as in traditional plane-wave DFT. Further tests and discussion of the FFT box technique will be published elsewhere [47].

Our next example involves the potential energy curve of the LiH molecule inside a large cubic simulation cell of side length $40a_0$.

Figure 3: Potential energy curves for LiH generated with the CASTEP plane-wave pseudopotential code and with our method for NGWF radii of $6.0a_0$ and $8.0a_0$ and for s-type PAOs with a radius of $6.0a_0$.

\scalebox{0.5}{\includegraphics*[0cm,1cm][25cm,18.5cm]{LiH_potential.eps} }

In Figure 3 the potential energy curve is shown as calculated by CASTEP and by our method with the same kinetic energy cutoff of 538eV. As we have used norm-conserving Troullier-Martins [52] pseudopotentials, this is a two electron system which we describe by one NGWF on each atom. It can be seen that when we use NGWFs with radii of $8.0a_0$, we have mE$_h$ agreement in total energies with the CASTEP results. Furthermore, the equilibrium bond length and vibrational frequency for this case differ from the CASTEP results by only -0.19% and 0.74% respectively. For the smaller radius of $6.0a_0$ the curve diverges from the CASTEP curve at large bond lengths. This is because the NGWF sphere overlap, and therefore the number of delta functions between the atoms, decreases more rapidly for the small radii as the atoms are pulled apart. Also shown in the same Figure is a curve that has been generated with our method but without optimisation of the NGWFs, which were kept constant and equal to the initial PAO guesses. This is equivalent to a tight-binding calculation with a minimal PAO basis. As can be seen, the total energies deviate significantly from the CASTEP result, as one would expect. The equilibrium bond length for this case differs by 3.34% from CASTEP, as compared to -1.24% for the $6.0a_0$ NGWF calculation. Thus, optimising the NGWFs improves the estimate of the bond length. The vibrational frequency obtained from the PAO case, however, differs by 3.51% from CASTEP, as compared to 5.33% for the $6.0a_0$ NGWF calculation. This, we believe, is an artefact of the localisation constraint imposed on the NGWFs and suggests that in fact localisation radii greater than $6.0a_0$ should be used in practice.

We now show convergence of the total energy with NGWF radius. For our tests we have used a silane (SiH$_4$) molecule with the same simulation cell and grid spacing as described above. A local Troullier-Martins [52] norm-conserving pseudopotential was used on the hydrogen atoms and a non-local one on the silicon atom. The number of NGWFs on each atom was as many as in the valence shells of the isolated atoms, i.e. one on hydrogen and four on silicon.

Figure 4: Total energy of a silane molecule calculated with our method for various NGWF radii (r$_c$).

\scalebox{0.5}{\includegraphics*{silane.eps} }

Figure 4 shows total energy results calculated for this system as a function of NGWF sphere radii. Convergence is uniform and to mE$_h$ accuracy by the time we get to a radius of $7.0a_{0}$. Such an NGWF radius should be adequate for practical calculations.

Here we also show that large qualitative changes occur to the shapes of the NGWFs during optimisation. In Figure 5 we show plots of isosurfaces of the NGWFs for an ethene molecule in a large simulation cell, before and after optimisation. The NGWF radius was $8.0a_0$ for all atoms.

Figure 5: Isosurfaces of NGWFs for the ethene molecule, before and after optimisation. The light grey surfaces are positive and the dark grey are negative. A drawing of the ethene molecule is superimposed on each NGWF in order to show its location with respect to the atoms.

  initial optimised
C 2p$_x$ \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,15.5cm]{ethylene_noC2s_C2px_fbl_BW.ps} } \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,15.5cm]{ethylene_noC2s_C2px_wan_BW.ps} }
C 2p$_y$ \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16cm]{ethylene_noC2s_C2py_fbl_BW.ps} } \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16cm]{ethylene_noC2s_C2py_wan_BW.ps} }
C 2p$_z$ \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16cm]{ethylene_noC2s_C2pz_fbl_BW.ps} } \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16cm]{ethylene_noC2s_C2pz_wan_BW.ps} }
H 1s \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16.5cm]{ethylene_noC2s_H1s_fbl_BW.ps} } \scalebox{0.5}{\includegraphics*[5.5cm,7cm][16cm,16.5cm]{ethylene_noC2s_H1s_wan_BW.ps} }

In particular, the carbon 2p$_x$ orbital, which is collinear with the C-C $\sigma$ bond, focusses more around this bond and gains two more lobes and nodes at the positions of the hydrogen atoms farthest from its carbon centre. The hydrogen functions, starting from 1s, obtain after optimisation a complicated shape that extends over the whole molecule and has nodal surfaces betwen the carbons and the rest of the hydrogens. The deep qualitative changes to the shapes of the NGWFs that occur during their optimisation with our method are obviously necessary for obtaining a plane-wave equivalent result. Our optimised NGWFs in general look nothing like the atomic orbitals they started from and are adjusted to their particular molecular environment. We therefore conclude that using the delta function basis set and performing all operations consistently with the plane-wave formalism is important for obtaining the systematic convergence that plane-waves have.

Our final example demonstrates the direct applicability of

Figure 6: A portion (Si$_8$H$_{16}$) of an infinite linear silane chain in a hexagonal simulation cell.

\scalebox{0.5}{\includegraphics*[3cm,3cm][19cm,20.5cm]{Si8H16_in_cell_BW.ps} }

our method to any lattice symmetry without any modification. This is a consequence of being consistent throughout with the plane-wave formalism. As we have shown for the calculation of the kinetic energy in this way [43], we also achieve better accuracy at no additional cost compared with a finite difference approach. In Figure 6 we show a portion (Si$_8$H$_{16}$) of an infinite linear silane chain inside a hexagonal simulation cell on which we have performed a total energy calculation at a kinetic energy cutoff of 183eV. The radii of the NGWFs were $6.0a_0$ on silicon and $5.0a_0$ on hydrogen. A total energy of -39.097E$_h$ was obtained when we optimised the density kernel only (with the NGWFs kept constant and equal to PAOs). When both the density kernel and the NGWFs were optimised, the energy lowered to -52.216E$_h$, which is another manifestation of the fact that both the density kernel and the NGWFs should be optimised in calculations with our method.


next up previous
Next: Conclusions Up: Nonorthogonal generalized Wannier function Previous: The FFT box technique
Peter D. Haynes 2002-10-31