|
|
Line 1: |
Line 1: |
| The [[Dielectric_properties|macroscopic dielectric function]], <math>\epsilon_{ij}(\omega)</math>, measures how a given dielectric medium reacts when subject to an external electric field. VASP possesses several methods to compute this function. Here we focus on a <math>O(N^2)</math> strategy, which can be useful when dealing with large systems. This method comes from the fact that the computation of the dielectric function can be reformulated as a time-propagation problem, instead of an eigenvalue equation. Below we present a brief description of the method, as well as the relevant {{TAG|INCAR}} variables. | | The [[Dielectric_properties|macroscopic dielectric function]], <math>\epsilon_{ij}(\omega)</math>, measures how a given dielectric medium reacts when subject to an external electric field. From <math>\epsilon_{ij}(\omega)</math> one can extract several optical properties such as absorption, optical conductivity, reflectance. However, it is important that the interacting electrons and holes are taken into account. This makes the evaluation of the macroscopic dielectric function more involved, since it goes beyond the single-particle level, working either at the [[Bethe-Salpeter equations|Bethe-Salpeter]] or [[Time-dependent density-functional theory calculations|time-dependent density-functional theory]] level. |
| {{NB|mind| This feature is only available for VASP.6 or higher!}}
| |
|
| |
|
| ==Macroscopic polarizability as a time-dependent integral== | | Within VASP, users can select two different methods for how <math>\epsilon_{ij}(\omega)</math> is computed. The first is based on the eigendecomposition of the electron-hole Hamiltonian, <math>H^\mathrm{exc}</math>. It allows for the evaluation of <math>\epsilon_{ij}(\omega)</math> by first obtaining the eigenvalues and eigenvectors of <math>H^\mathrm{exc}</math> and it is based on the [[Bethe-Salpeter equations|Bethe-Salpeter equation]] or the [[Time-dependent density-functional theory calculations|Casida equation]]. The second method transforms the mathematical expression of <math>\epsilon_{ij}(\omega)</math> into a time-dependent integral. By propagating in time the dipolar moments and then applying a Fourier transform, it can compute <math>\epsilon_{ij}(\omega)</math>. |
| The starting point is that one can re-write <math>\epsilon_{ij}(\omega)</math> as a time-dependent integral{{cite|schmidt:prb:2003}} | | |
| | The advantage of the later method in comparison to the former is related to the cost, with the time-dependent integral being <math>O(N^2)</math>, while the eigendecomposition has a cost of <math>O(N^3)</math>, where <math>N</math> is the rank of <math>H^\mathrm{exc}</math>. This means that for very large numbers of bands or k-points, the time-dependent formalism is cheaper than the eigendecomposition method. |
| | |
| | Below we present a brief description of the method, from its theoretical support to how calculations should be performed, with the relevant approximations needed in the two-particle Hamiltonian. |
| | |
| | ==The macroscopic-dielectric function as a time-dependent integral== |
| | The starting point is that one can re-write <math>\epsilon_{ij}(\omega)</math> as a time-dependent integral{{cite|schmidt:prb:2003}}. It starts from its expression, given by |
|
| |
|
| ::<math> | | ::<math> |
| \epsilon_{ij}(\omega)=\delta_{ij}-\frac{4\pi e^2}{\Omega}\int_0^{\infty} \mathrm{d} t | | \epsilon^M(\omega)=1+\frac{4 \pi}{\Omega_0} \sum_\lambda\left|\sum_{c v \mathbf{k}} \mu_{c v \mathbf{k}} A_{c v \mathbf{k}}^\lambda\right|^2\left[\frac{1}{\omega+E_\lambda+\mathrm{i} \eta}-\frac{1}{\omega-E_\lambda+\mathrm{i} \eta}\right] |
| \sum_{c,v,\mathbf{k}}\left(\langle\mu^j_{cv\mathbf{k}}| \xi^i_{cv\mathbf{k}}(t)\rangle+ \mathrm{c.c.}\right) e^{-\mathrm i(\omega-\mathrm i \delta) t} | |
| </math>, | | </math>, |
|
| |
|
| where <math>\mu_{v c \mathbf{k}}^j=\frac{\left\langle c \mathbf{k}\left|v_j\right| v \mathbf{k}\right\rangle}{\varepsilon_c(\mathbf{k})-\varepsilon_v(\mathbf{k})}</math> is a vector of all dipole moments. The fundamental aspect behind this transformation is that the new, time-dependent vector <math>\left.\mid \xi^j(t)\right\rangle</math> follows the equation | | where <math>\mu_{v c \mathbf{k}}^j=\frac{\left\langle c \mathbf{k}\left|v_j\right| v \mathbf{k}\right\rangle}{\varepsilon_c(\mathbf{k})-\varepsilon_v(\mathbf{k})}</math> is the dipolar moment associated to the the conduction <math>c</math> and valence band <math>v</math>, and k-point <math>k</math>, <math>\lambda</math> is the index of the eigenstate of <math>H^\mathrm{exc}</math>, with <math>A^\lambda</math> and <math>E_\lambda</math> being the associated eigenvector and eigenvalue. This equation can be brought into operational form, |
|
| |
|
| ::<math> | | ::<math> |
| \mathrm i \frac{\mathrm d}{\mathrm d t}\left|\xi^j(t)\right\rangle=\hat{H}(t)\left|\xi^j(t)\right\rangle, | | \epsilon^M(\omega)=1+\frac{4 \pi}{\Omega_0}\left\langle\mu\left|\left[\frac{1}{\omega+\mathrm{i} \eta+\hat{H}^{\mathrm{exc}}}-\frac{1}{\omega+\mathrm{i} \eta-\hat{H}^{\mathrm{exc}}}\right]\right| \mu\right\rangle |
| </math> | | </math> |
| | by using the spectral decomposition <math>\left[\hat{H}^{\mathrm{exc}}-\omega\right]^{-1}=\sum_\lambda \frac{\left|A_\lambda\right\rangle\left\langle A_\lambda\right|}{E_\lambda-\omega}</math>. Then, one can bring the new expression of <math>\epsilon(\omega)</math> into a time-dependent integral, by using |
|
| |
|
| with the initial vector elements given by <math>\left|\xi^j(0)\right\rangle=\left|\mu^j\right\rangle</math>. It is worthy to note that <math>H(t)</math> is split into two parts, <math>H(t) = H_0 + V(t)</math>, where <math>H_0</math> is the ground-state Hamiltonian, and <math>V(t)</math> is the time-dependent perturbation.
| | ::<math> |
| | \frac{1}{\omega+\mathrm{i} \eta-\hat{H}^{\mathrm{exc}}}|\mu\rangle=-\mathrm{i} \int_0^{\infty} e^{-\mathrm{i}\left(\omega-\hat{H}^{\mathrm{exc}}+\mathrm{i} \eta\right) t}|\mu\rangle=-\mathrm{i} \int_0^{\infty} e^{-\mathrm{i}(\omega+\mathrm{i} \eta) t} e^{\mathrm{i} \hat{H}^{\mathrm{exc}}t}|\mu\rangle |
| | </math>, |
| | and recognising that <math>e^{\mathrm{i} \hat{H}^{\mathrm{exc}}t}|\mu\rangle = |\xi(t)\rangle</math> is the exponential form of a time-dependent equation. These considerations allow the expression of <math>\epsilon_{ij}(\omega)</math> to be written as |
|
| |
|
| To probe the dielectric response, <math>V(t)</math> is set as proportional to a delta-pulse{{cite|kubo:jpsj:1957}}, i.e. <math>V(t) = \lambda\mathbf r\cdot\mathbf D\delta(t)</math>{{cite|sander:jcp:2017}}. With this, all possible transitions between the conduction and valence states included in the INCAR are probed during time-propagation. Control over which valence and conduction states are added in time-dependent run is done with the {{TAG|NBANDSO}} and {{TAG|NBANDSV}} flags, respectively.
| | ::<math> |
| | \epsilon_{ij}(\omega)=\delta_{ij}-\frac{4\pi e^2}{\Omega}\int_0^{\infty} \mathrm{d} t |
| | \sum_{c,v,\mathbf{k}}\left(\langle\mu^j_{cv\mathbf{k}}| \xi^i_{cv\mathbf{k}}(t)\rangle+ \mathrm{c.c.}\right) e^{-\mathrm i(\omega-\mathrm i \delta) t} |
| | </math>, |
|
| |
|
| By propagating the projection of <math>\mu^i</math> onto <math>\xi^j(t)</math> at each instant, <math>c_{cv\mathbf k}^{ij}(t)</math>, VASP can compute the dielectric response along different directions (<math>i,j=x,y,z</math>) by evaluating the Fourier transform of each time-dependent coefficient.
| | The fundamental aspect behind this transformation is that the new, time-dependent vector <math>\left.\mid \xi^j(t)\right\rangle</math> follows the equation |
| {{NB|mind| Since the problem is now a time-dependent equation, there is no computation of eigenvalues and eigenvectors of the two-particle Hamiltonian, unlike what happens in the [[Bethe-Salpeter-equations_calculations|Bethe-Salpeter formulation]].}}
| |
|
| |
|
| With the variable {{TAG|IEPSILON}} users can control the Cartesian direction along which the Dirac delta pulse is applied. The default value, {{TAG|IEPSILON}} = 4, performs three independent calculations for an electric field in x, y, and z direction, and thus is the most expensive. The choice of direction is system dependent: while for cubic systems it is enough to compute <math>\epsilon_{xx}(\omega)</math>, for 2D systems (eg. monolayer hexagonal boron nitride) it is advisable to compute both <math>\epsilon_{xx}(\omega)</math> and <math>\epsilon_{yy}(\omega)</math> and then analyse their average. This assuming that the <math>(001)</math> direction is perpendicular to the plane in which the atoms are laid.
| |
|
| |
| Users should note that a sufficient large number of such states must be included in a preceding ground-state calculation, where all the energy levels and dipole moments are computed. Note, however, that unoccupied orbitals are not propagated, which saves computational time. Also, if the interest is to compare time-evolution results with data coming from BSE/TDDFT calculations, users should set the same values for {{TAG|NBANDSO}} and {{TAG|NBANDSV}} in both calculations.
| |
|
| |
| Finally, to perform this calculation users must set {{TAG|ALGO}}={{TAG|TIMEEV}} in the INCAR file. A small example for bulk silicon is provided in the last section of this page.
| |
|
| |
| ==Levels of approximation for the exchange-correlation potential==
| |
|
| |
| The interaction between electrons and holes can be controlled at three levels of approximation, depending on the input provided by the user. They are controlled by the {{TAG|INCAR}} variables, {{TAG|LADDER}}, {{TAG|LHARTREE}}, and {{TAG|LFXC}}. In this section we provide a brief description and some advice on when to use them.
| |
|
| |
| ===Independent particle approximation (IPA)===
| |
| This is the default setting of VASP, when all three tags are missing. Users can use this setup to perform checks on whether or not all necessary files are present and correct, by comparing the obtained spectrum with that computed in the DFT step with {{TAG|LOPTICS}}=.TRUE.
| |
|
| |
| ===LFXC: activation of the exchange-correlation kernel===
| |
| Setting {{TAG|LFXC}}=.TRUE. includes the local exchange-correlation kernel, <math>f_\mathrm{xc}</math> in the time-propagation
| |
| ::<math> | | ::<math> |
| f_{\mathrm{xc}}^{\text {loc }}\left(\mathbf{r}, \mathbf{r}^{\prime}\right)=\frac{\delta^2\left\{E_{\mathrm{c}}^{\mathrm{DFT}}+\left(1-c_{\mathrm{x}}\right) E_{\mathrm{x}}^{\mathrm{DFT}}\right\}}{\delta \rho(\mathbf{r}) \delta \rho\left(\mathbf{r}^{\prime}\right)},
| | \mathrm i \frac{\mathrm d}{\mathrm d t}\left|\xi^j(t)\right\rangle=\hat{H}(t)\left|\xi^j(t)\right\rangle, |
| </math> | | </math> |
| where <math>c_X</math> controls the fraction of the exchange energy functional that is included in the kernel (see {{TAG|AEXX}}). This allows users to perform time-dependent calculations using hybrid functionals.
| |
|
| |
| The accuracy of local exchange-correlation kernels when compared to experimental data if often limited to molecules up to around 100-200 atoms. As these kernels often lack the long range component that goes as <math>-1/q^2</math> (where <math>q</math> is the momentum difference between the electron and the hole), they fail to properly reproduce the binding energies of electron-hole pairs in periodic systems or very large molecules.
| |
| {{NB|mind|If none of the three tags are present, VASP will perform a TD-DFT calculation by default!}}
| |
|
| |
| ===LADDER: activation of the screened exchange interaction===
| |
| By setting {{TAG|LADDER}}=.TRUE., users can turn on the inclusion of the ladder diagrams, which account for what are usually called excitonic effects. These come from the screened exchange interaction potential, <math>W(\mathbf r,\mathbf r';\omega)</math> and have the right long-range behaviour, allowing for the formation of bound electron-hole pairs in bulk materials.
| |
|
| |
| At the present, the screened interaction has to be computed from a [[Improving_the_dielectric_function#Model-BSE|model dielectric function]]. Users should add the following tags to the {{TAG|INCAR}} file
| |
|
| |
| {{TAG|LHFCALC}} = .TRUE.
| |
| {{TAG|LMODELHF}} = .TRUE.
| |
| {{TAG|HFSCREEN}} = 1.26
| |
| {{TAG|AEXX}} = 0.1
| |
|
| |
| and it is also advisable to check the recommended values of {{TAG|HFSCREEN}} and {{TAG|AEXX}} for specific materials on the page of each respective tag.
| |
|
| |
| For small molecules the inclusion of ladder diagrams is often unnecessary, but the long range interaction becomes important for large molecules.
| |
|
| |
| ===LHARTREE: activation of the unscreened Coulomb interaction===
| |
| With the flag {{TAG|LHARTREE}}=.TRUE. users can activate the inclusion of direct interaction terms in the Hamiltonian, also known as the bubble diagrams from many-body perturbation theory (MBPT). This is the equivalent of running a BSE calculation in the random phase approximation. Note that at the end, the dielectric function reported in the output files is the macroscopic dielectric function, where no contributions from local fields (i.e. terms with finite <math>\mathbf G</math>) are included.
| |
|
| |
| ==Choice of time-step==
| |
| By default, the number of time steps is chosen automatically by VASP and and set to 20000. Alternatively, the number of time steps can be set directly by the tag {{TAG|NELM}}. In this case, the user-defined number of steps needs to be large than 100. Otherwise, the value of {{TAG|NELM}} will be discarded and VASP will revert to the default value.
| |
|
| |
| ==Summary of required INCAR flags==
| |
| ===Run-level===
| |
| {{TAG|ALGO}} = TIMEEV - sets the calculation type to a time-evolution run.
| |
|
| |
| ===Choice of bands===
| |
| {{TAG|NBANDSO}} - number of occupied (valence) bands to be included.
| |
|
| |
| {{TAG|NBANDSV}} - number of virtual (conduction) bands to be included.
| |
|
| |
| ===Approximations to the potential===
| |
| {{TAG|LADDER}} - include or exclude screened exchange interaction (i.e. ladder diagrames) in the Hamiltonian.
| |
|
| |
| {{TAG|LHARTREE}} - include or exclude the Hartree exchange terms (i.e. bubble diagrams) in the Hamiltonian.
| |
|
| |
| {{TAG|LFXC}} - include or exclude the <math>f_\mathrm{xc}</math> kernel in the time-propagation.
| |
|
| |
| {{TAG|LMODELHF}} - uses a range separated model dielectric function to compute the screened interaction. Users must specify two range separation parameters, {{TAG|HFSCREEN}} and {{TAG|AEXX}}.
| |
|
| |
| {{TAG|HFSCREEN}} - controls the real-space range separation parameter.
| |
|
| |
| {{TAG|AEXX}} - controls the fraction of exchange functional that is included in the <math>f_\mathrm{xc}</math>, or in the case of the model dielectric function the value of <math>\epsilon^{-1}_\infty</math>.
| |
|
| |
| ===Time-propagation and spectrum===
| |
| {{TAG|CSHIFT}} - controls the broadening of the peaks in the spectra of <math>\epsilon_{ij}(\omega)</math>, while also automatically setting up the maximum time used in the time-propagation algorithm.
| |
|
| |
| {{TAG|IEPSILON}} - sets up the directions along which the dielectric function is computed.
| |
|
| |
| {{TAG|NELM}} - sets the number of steps to be used in time-propagation. Users must set this number to be greater than 100 if they want to avoid the default value (20000).
| |
|
| |
| {{TAG|OMEGAMAX}} - maximum frequency (in eV) for which the dielectric function is computed.
| |
|
| |
| ==TIMEEV versus TDHF==
| |
| VASP provides another algorithm with which users can compute the macroscopic dielectric function, {{TAG|TDHF}}. Here however, the problem is formulated in terms of an eigenvalue decomposition via the Casida equation. Both approaches are equivalent, although {{TAG|TIMEEV}} is faster for adiabatic LDA and PBE, while {{TAG|TDHF}} is faster for hybrid functionals.
| |
|
| |
| == Example ==
| |
|
| |
| A typical calculation requires two steps. First, a ground state calculation:
| |
|
| |
| {{TAGBL|SYSTEM}} = Si
| |
| {{TAGBL|NBANDS}} = 12 ! even 8 bands suffice for Si
| |
| {{TAGBL|ISMEAR}} = 0 ; {{TAGBL|SIGMA}} = 0.05
| |
| {{TAGBL|ALGO}} = N
| |
| {{TAGBL|LOPTICS}} = .TRUE.
| |
| {{TAGBL|KPAR}} = 4 ! assuming we run on 4 cores, this will be the fastest
| |
|
| |
|
| | with the initial vector elements given by <math>\left|\xi^j(0)\right\rangle=\left|\mu^j\right\rangle</math>. It is worthy to note that <math>H(t)</math> is split into two parts, <math>H(t) = H_0 + V(t)</math>, where <math>H_0</math> is the ground-state Hamiltonian, and <math>V(t)</math> is the time-dependent perturbation. |
|
| |
|
| Second, the actual time propagation:
| | ==The delta-like perturbation== |
| | explain how the system is perturbed with the delta-potential |
|
| |
|
| {{TAGBL|SYSTEM}} = Si
| | ==The many-body terms in the hamiltonian== |
| {{TAGBL|NBANDS}} = 12 ! even 8 bands suffice for Si
| | ===Independent-particle approximation=== |
| {{TAGBL|ISMEAR}} = 0 ; {{TAGBL|SIGMA}} = 0.05
| | ===Hartree exchange potential=== |
| {{TAGBL|ALGO}} = TIMEEV
| | ===Screened two-particle interaction=== |
| {{TAGBL|IEPSILON}} = 1 ! cubic system, so response in x direction suffices
| | ====Exchange-correlation effects from time-dependent density functional theory==== |
| {{TAGBL|NBANDSO}} = 4 ; {{TAGBL|NBANDSV}} = 8 ; {{TAGBL|CSHIFT}} = 0.1
| | ====Ladder diagrams from many-body perturbation theory==== |
| {{TAGBL|KPAR}} = 4 ! assuming we run on 4 cores, this will be the fastest
| | explain what different components are included in H (LFXC,LHARTREE,LADDER, or none) |
|
| |
|
| In this case, {{TAG|OMEGAMAX}} is set automatically to the maximal transition energy (about 25 eV in this example). Reducing the number of considered transitions, and thus reducing {{TAG|OMEGAMAX}} will increase both the duration of time steps and their number.
| | ==Step-by-step instructions== |
|
| | ===Step 1: Ground state with extra empty states=== |
| For standard DFT calculations, the time-propagation code is so fast that very dense k-point grids can often be used.
| | ===Step 2: Time-evolution run=== |
| | ====Setting up the time-step==== |
| | ====Choosing the direction of perturbation==== |
|
| |
|
| == Related variables and articles == | | ==Comparison to other methods== |
| {{TAG|ALGO}},
| | ===Bethe-Salpeter equation=== |
| {{TAG|CSHIFT}},
| | ===Casida equation=== |
| {{TAG|LHARTREE}},
| |
| {{TAG|LFXC}},
| |
| {{TAG|NBANDSV}},
| |
| {{TAG|NBANDSO}},
| |
| {{TAG|OMEGAMAX}}
| |
|
| |
|
| == References == | | ==Related tags and articles== |
|
| |
|
| | ==References== |
|
| |
|
| <!---[[Category:Linear response]] | | <!---[[Category:Linear response]] |
The macroscopic dielectric function,
, measures how a given dielectric medium reacts when subject to an external electric field. From
one can extract several optical properties such as absorption, optical conductivity, reflectance. However, it is important that the interacting electrons and holes are taken into account. This makes the evaluation of the macroscopic dielectric function more involved, since it goes beyond the single-particle level, working either at the Bethe-Salpeter or time-dependent density-functional theory level.
Within VASP, users can select two different methods for how
is computed. The first is based on the eigendecomposition of the electron-hole Hamiltonian,
. It allows for the evaluation of
by first obtaining the eigenvalues and eigenvectors of
and it is based on the Bethe-Salpeter equation or the Casida equation. The second method transforms the mathematical expression of
into a time-dependent integral. By propagating in time the dipolar moments and then applying a Fourier transform, it can compute
.
The advantage of the later method in comparison to the former is related to the cost, with the time-dependent integral being
, while the eigendecomposition has a cost of
, where
is the rank of
. This means that for very large numbers of bands or k-points, the time-dependent formalism is cheaper than the eigendecomposition method.
Below we present a brief description of the method, from its theoretical support to how calculations should be performed, with the relevant approximations needed in the two-particle Hamiltonian.
The macroscopic-dielectric function as a time-dependent integral
The starting point is that one can re-write
as a time-dependent integral[1]. It starts from its expression, given by
,
where
is the dipolar moment associated to the the conduction
and valence band
, and k-point
,
is the index of the eigenstate of
, with
and
being the associated eigenvector and eigenvalue. This equation can be brought into operational form,
![{\displaystyle
\epsilon^M(\omega)=1+\frac{4 \pi}{\Omega_0}\left\langle\mu\left|\left[\frac{1}{\omega+\mathrm{i} \eta+\hat{H}^{\mathrm{exc}}}-\frac{1}{\omega+\mathrm{i} \eta-\hat{H}^{\mathrm{exc}}}\right]\right| \mu\right\rangle
}](/wiki/index.php?title=Special:MathShowImage&hash=c1a36161307730e99c2fa46831ce83c9&mode=mathml)
by using the spectral decomposition
. Then, one can bring the new expression of
into a time-dependent integral, by using
,
and recognising that
is the exponential form of a time-dependent equation. These considerations allow the expression of
to be written as
,
The fundamental aspect behind this transformation is that the new, time-dependent vector
follows the equation
![{\displaystyle
\mathrm i \frac{\mathrm d}{\mathrm d t}\left|\xi^j(t)\right\rangle=\hat{H}(t)\left|\xi^j(t)\right\rangle,
}](/wiki/index.php?title=Special:MathShowImage&hash=074ab8aca03155dac09394d03ce7f584&mode=mathml)
with the initial vector elements given by
. It is worthy to note that
is split into two parts,
, where
is the ground-state Hamiltonian, and
is the time-dependent perturbation.
The delta-like perturbation
explain how the system is perturbed with the delta-potential
The many-body terms in the hamiltonian
Independent-particle approximation
Hartree exchange potential
Screened two-particle interaction
Exchange-correlation effects from time-dependent density functional theory
Ladder diagrams from many-body perturbation theory
explain what different components are included in H (LFXC,LHARTREE,LADDER, or none)
Step-by-step instructions
Step 2: Time-evolution run
Setting up the time-step
Choosing the direction of perturbation
Comparison to other methods
Bethe-Salpeter equation
Casida equation
Related tags and articles
References