next up previous contents
Next: 3.5 Summary Up: 3. Quantum Monte Carlo Previous: 3.3 Variational Monte Carlo   Contents

Subsections

3.4 Diffusion Monte Carlo

Diffusion Monte Carlo (DMC) is a projector or Green's function based method for solving for the ground state of the many-body Schrödinger equation. In principle the DMC method is exact, although in practice, several well-controlled approximations must be introduced for calculations to remain tractable.

3.4.1 Outline of method

The DMC method is based on rewriting the Schrödinger equation in imaginary time, $\tau = it$. The imaginary time Schrödinger equation is:
\begin{displaymath}
{\partial \vert\psi\rangle \over \partial\tau} = -\hat H \vert\psi\rangle \;\;\;.
\end{displaymath} (3.25)

The state $\vert\psi\rangle$ is expanded in eigenstates of the Hamiltonian
\begin{displaymath}
\vert\psi\rangle=\sum_{i=0}^{\infty} c_{i} \vert\phi_i\rangle \;\;\;,
\end{displaymath} (3.26)

where
\begin{displaymath}
\hat{H}\vert\phi_i\rangle = \epsilon_i \vert\phi_i\rangle \;\;\;.
\end{displaymath} (3.27)

A formal solution of the imaginary time Schrödinger equation is
\begin{displaymath}
\vert\psi(\tau_{1}+\delta\tau)\rangle
=e^{-\hat{H}\delta\tau}\vert\psi(\tau_1)\rangle \;\;\;,
\end{displaymath} (3.28)

where the state $\vert\psi\rangle$ evolves from imaginary time $\tau_1$ to a later time $\tau_1+\delta\tau$. If the initial state $\vert\psi(\tau_1)\rangle$ is expanded in energy ordered eigenstates, following equation 3.26, then
\begin{displaymath}
\vert\psi(\delta\tau)\rangle=
\sum_{i=0}^{\infty}c_{i}e^{-\epsilon_{i}\delta\tau}\vert\phi_{i}\rangle \;\;\;.
\end{displaymath} (3.29)

Hence any initial state, $\vert\psi\rangle$, that is not orthogonal to the ground state, $\vert\phi_{0}\rangle$, will evolve to the ground state in the long time limit:
\begin{displaymath}
\lim_{\tau\to\infty} \vert\psi(\tau)\rangle = c_0 e^{-\epsilon_{0} \tau}
\vert\phi_{0}\rangle \;\;\;.
\end{displaymath} (3.30)

This derivation shares many formal similarities with that given for the variational principle in section 3.3.1. However in the DMC method the imaginary time evolution results in excited states decaying exponentially fast, whereas in the VMC method any excited state contributions remain and contribute to the VMC energy.

The DMC method is a realisation of the above derivation in position space. equation 3.30 becomes

\begin{displaymath}
\lim_{\tau\to\infty} \psi({\bf R},\tau) = c_0 e^{-\epsilon_{0} \tau}
\phi_{0}({\bf R}) \;\;\;.
\end{displaymath} (3.31)

By introducing a constant offset to the energy, $E_{T}=\epsilon_0$, the long-time limit of equation 3.31 can be kept finite. If the Hamiltonian is separated into the kinetic energy and potential terms, the imaginary time Schrödinger equation, equation 3.25, takes on a form similar to a diffusion equation:
\begin{displaymath}
-\frac{\partial \psi ({\bf R}, \tau )}{\partial\tau}=
\left[...
... R},\tau)\right]
+(V({\bf R})-E_{T})\psi({\bf R},\tau) \;\;\;.
\end{displaymath} (3.32)

This equation is a diffusion equation where $\psi({\bf R})$ may be interpreted as the density of diffusing particles (or ``walkers''), and the term $(V({\bf R})-E_{T})$ is a rate term describing a potential-dependent increase or decrease in the particle density. This is shown schematically in one-dimension for a simple potential in figure 3.1.

Figure 3.1: The evolution of walkers in the DMC method. A single particle in confined in a simple potential well, $V(x)$. An initially even distribution of walkers (circles) is propagated in imaginary time, $\tau$. The distribution gradually evolves by a process of diffusion and branching to a distribution representative of the ground state wavefunction, $\psi(x)$. Note that where the potential energy is low, walkers tend to branch giving a higher density of walkers, and where the potential energy is high, walkers tend to be removed.
\includegraphics [width=10cm]{Figures/branch.eps}

The above equation may be transformed into a form suitable for Monte Carlo methods, [23,27] but this leads to a very inefficient algorithm. The potential $V({\bf R})$ is unbounded in coulombic systems and hence the rate term, $V({\bf R})-E_{T}$, can diverge. Large fluctuations in the particle density then result and give impractically large statistical errors. These fluctuations may be substantially reduced by the incorporation of importance sampling in the algorithm.

The above method also lacks the important constraint that for electronic systems the solution should be antisymmetric with respect to the exchange of any two electrons, i.e. is fermionic. This constraint may be imposed by assuming a form for the nodal surface of the fermionic wavefunction, but at the expense of obtaining only a variational solution to the exact ground state energy. This method of constraint, the fixed-node approximation, and other methods of enforcing fermionic anti-symmetry are discussed in section 3.4.4.

3.4.2 Importance sampling

Importance sampling is essential for DMC methods, if the simulation is to be efficient. [27] A trial or guiding wavefunction, $\Psi_{G}({\bf R})$, which closely approximates the ground state wavefunction is introduced. The derivation of Reynolds et. al.[28,23] is followed in the remainder of this section.

A new distribution is defined

\begin{displaymath}
f({\bf R},\tau )=\psi_{G}({\bf R})\psi({\bf R},\tau) \;\;\;,
\end{displaymath} (3.33)

which is also a solution of the Schrödinger equation when $\psi({\bf R},\tau)$ is a solution. The equation for non-importance sampled DMC, equation 3.32, is consequently modified:
\begin{displaymath}
-\frac{\partial f({\bf R},\tau)}{\partial\tau}=
\left[\sum_{...
...R},\tau)\right]
+(E_{L}({\bf R})-E_{T})f({\bf R},\tau) \;\;\;.
\end{displaymath} (3.34)

$E_{T}$ is a ``trial energy'' introduced to maintain normalisation of the projected solution at large $\tau$. The term
\begin{displaymath}
{\bf F}({\bf R})=\frac{\nabla\psi({\bf R})}{\psi({\bf R})}\;\;\;,
\end{displaymath} (3.35)

is commonly referred to as the ``quantum force''[29]. The local energy, $E_{L}({\bf R})$, equation 3.17 is computed with respect to the guiding wavefunction $\psi_{G}({\bf R})$.

The right hand side of the importance sampled DMC equation (equation 3.34) consists, from left to right, of diffusion, drift and rate terms. The problematic potential dependent rate term of the non-importance sampled method is replaced by a term dependent on the difference between the local energy of the guiding wavefunction and the trial energy. The trial energy is initially chosen to be the VMC energy of the guiding wavefunction, and is updated as the simulation progresses. Use of an optimised guiding function minimises the difference between the local and trial energies, and hence minimises fluctuations in the distribution $f$. A wavefunction optimised using VMC is ideal for this purpose, and in practice VMC provides the best method for obtaining wavefunctions that accurately approximate ground state wavefunctions locally. The guiding wavefunction may be also constructed to minimise the number of divergences in $E_{L}({\bf R})$, unlike the non-importance sampled method where divergences in the coulomb interactions are always present.


3.4.3 Transformation to integral form

To be of use, the importance sampled DMC equation equation 3.34 must be transformed into a form suitable for Monte Carlo integration. The transformation is more complex than for VMC, which simply required the insertion of the factor $\vert\psi({\bf R})\vert^2/\int \vert\psi({\bf R})\vert^2 d{\bf R}$ into the conventional formulas for quantum mechanical expectation values.

A Green's function, $G({\bf R}^\prime,{\bf R};\tau)$, that is a solution of equation 3.34 is desired, i.e, a spatial representation of the imaginary time propagator, $\exp[-\tau(\hat{H}-E_{T})]$:


\begin{displaymath}
f({\bf R}^\prime,\tau_{0}+\tau)=\int G({\bf R}^\prime,{\bf R};\tau)f({\bf R},\tau_{0})d{\bf R}
\end{displaymath} (3.36)

The Green's function may be approximated to second order in $\tau$ by factorising the propagator into branching and diffusion parts: [28,23]

\begin{displaymath}
G({\bf R}^\prime,{\bf R};\tau) \approx \tilde{G}({\bf R}^\prime,{\bf R};\tau)\equiv \tilde{G}_{b}\tilde{G}_{d} \;\;\;.
\end{displaymath} (3.37)

The branching part of the Green's function, $\tilde{G}_{b}$, is a function of the ``surplus local energy'':

\begin{displaymath}
\tilde{G}_{b}({\bf R}^\prime,{\bf R};\tau) =
\exp[(-\frac{1}{2}(E_{L}({\bf R}^\prime)+E_{L}({\bf R}))-E_{T})\tau] \;\;\;.
\end{displaymath} (3.38)

The diffusion part is obtained by assuming that the quantum force, ${\bf F}$, is constant over the move ${\bf R} \rightarrow {\bf R}^\prime$:

\begin{displaymath}
\tilde{G}_{d}({\bf R}^\prime,{\bf R};\tau) =
(2\pi \tau)^{-3...
...f R}^\prime-\tau {\bf F}({\bf R}^\prime)/2)^{2}/{2\tau}] \;\;.
\end{displaymath} (3.39)

Although accurate to second order in $\tau$, this approximation is significant in all-electron calculations, where ${\bf F}$ varies rapidly in core regions. Additionally, this component of the short-time approximation Green's function, $\tilde{G}$, is not Hermitian. ``Detailed balance'' is imposed with the conventional Metropolis probability
\begin{displaymath}
p=\mathrm{min}(1,W({\bf R}^\prime,{\bf R};\tau)) \;\;\;,
\end{displaymath} (3.40)

where
\begin{displaymath}
W({\bf R}^\prime,{\bf R};\tau)=\frac{\vert\Psi_{G}({\bf R}^\...
...\bf R})\vert^2 \tilde{G}({\bf R},{\bf R}^\prime;\tau)} \;\;\;,
\end{displaymath} (3.41)


3.4.4 The fixed-node approximation

The DMC method discussed so far lacks a constraint crucial for the simulation of fermionic systems. In general, the ground state wavefunction of a fermionic system has nodes and consequently regions of positive and negative sign. These regions are essential for a wavefunction to be anti-symmetric with respect to the interchange of any two fermions. The methods discussed so far rely on a probability or particle distribution which by definition is positive definite. If the constraint of antisymmetry is not imposed, the application of the DMC method results in a bosonic solution of lower energy than the true fermionic ground state energy being obtained. The most successful attempt to address to this ``nodal problem'' is the fixed-node approximation. It is used in all of the DMC calculations presented in this thesis.

The fixed-node approximation, due to Anderson, [30] was the first solution of the nodal problem. The approximation consists of imposing the nodes of a reference wavefunction. This approximation is variational: the fixed-node DMC energy is an upper bound to the exact ground state energy.[31] In Anderson's original implementation, walkers attempting to cross nodes were deleted. This has been shown to introduce a bias that is linear in the time step, [29] and all current implementations reject moves that cross nodes. A bias of second order in the time step results, which is the same order as the short-time approximation to the Green's function.

The accuracy of the fixed-node approximation is wholly dependent upon the nodes of the reference wavefunction. In practice the nodes of an HF or DFT based wavefunction are found to be very accurate, giving energies well below those of the best VMC wavefunctions. This is fortunate as currently there are no general methods for systematically improving the DMC energy that scale polynomially with the size of the system.

3.4.5 Improving the fixed-node energy

Two methods for systematically improving the fixed-node energy exist: A third method of improving the nodes exists. Calculations by Kwon[32,33] have shown that small improvements to DMC energies may be obtained for the 2D and 3D electron gases, at high densities, by replacing the positions of the electrons in the wavefunctions by ``quasiparticle coordinates'' [34] parameterised in terms of the positions of (in principle) all the electrons:
\begin{displaymath}
{\bf x}_{i}={\bf r}_{i}+\sum_{j\neq i}^{N}\eta(r_{ij})({\bf r}_{i}-{\bf r}_{j}) \;\;\;.
\end{displaymath} (3.42)

$\eta$ is a correlation or backflow[34] function containing variable parameters. Presently it is unclear whether the use of quasiparticle coordinates will be of use in real electronic systems. These methods carry considerable overhead, as the full wavefunction must be calculated when a single electron is moved. Although these methods may prove useful in small systems, the use of quasiparticle coordinates in large systems is likely to restricted, as the cost of moving all of the electrons is now $O(N^4)$ due to the additional cost of wavefunction recalculation (see chapter 4).

Additionally, these methods are perturbative in the sense that they modify an underlying reference nodal set. Although the use of quasiparticle coordinates may result in small improvements in real electronic systems, where good quality nodal sets are already available, their use cannot be considered a solution to the nodal problem. They are however, one of the few ideas present in the literature that scales reasonably well with system size.


3.4.6 The DMC algorithm

The DMC algorithm used for most of the calculations in this thesis is given below. It is similar to those of Refs. [28,29,23]. As with VMC there are distinct equilibration and accumulation phases, but both are almost identical as energies are required at every move in the calculation in the evaluation of the configuration branching factor.

$\Psi({\bf r}^\prime)$ is used to indicate the wavefunction with a single electron moved to point ${\bf r}^\prime$ and all other electrons unmoved:

  1. Initialise an ensemble of $N_{c}$ configurations, which should be uncorrelated and distributed according to the probability density of the guiding function $\vert\Psi_G\vert^2$. Initialise the trial energy $E_{T}$ to the average VMC energy of the ensemble.
  2. For every configuration $j$:
    1. For every electron $i$ in the current configuration:
      1. Propose a move according to


        \begin{displaymath}
{\bf r}_{i}^\prime={\bf r}_{i}+\tau{\bf F}({\bf r}_{i})+{\bf\eta} \;\;\;,
\end{displaymath} (3.43)

        where ${\bf F}$ is the ``quantum force'' of equation 3.35, and ${\bf\eta}$ is a Gaussian random vector of mean zero and variance $\tau$ per Cartesian component.
      2. Apply the fixed-node approximation, checking that $\Psi_{G}({\bf r}^\prime)$ is of the same sign as $\Psi_{G}({\bf r})$. If a node has been crossed, reject the move of electron $i$ and consider the next electron.
      3. Compute the weight for the move
        \begin{displaymath}
W({\bf r}^\prime,{\bf r})=\frac{\vert\Psi_{G}({\bf r}^\prime...
...\bf r})\vert^2 \tilde{G}({\bf r},{\bf r}^\prime;\tau)} \;\;\;,
\end{displaymath} (3.44)

        where $\tilde{G}$ is the short-time approximation to the Green's function (section 3.4.3).
      4. Accept the move according to the Metropolis probability $\mathrm{min}(1,W({\bf r}^\prime,{\bf r}))$

    2. Compute the branching factor for the moved configuration $j$:


      \begin{displaymath}
P_{B} = \exp\left(-\tau
(\frac{1}{2}\left[E_{L}({\bf R}^\prime)+E_{L}({\bf R})\right]-E_{T})\right)
\end{displaymath} (3.45)

    3. Accumulate the energy and any observables weighted by the branching factor $P_{B}$
    4. Make $\mathrm{int}(P_{B}+u)$ copies of the configuration, where $u$ is a uniformly distribution random number of $[0,1]$.
  3. Repeat the configuration moves for a number of steps (a block), typically $O(100-1000)$.
  4. Update the trial energy $E_{T}$, to bring it closer to that of the current ensemble, by using the average energy for the last block.
  5. Renormalise the number of walkers in the ensemble to the target number $N_c$ by randomly creating or deleting walkers.
  6. Repeat until sufficient data is obtained.

During equilibration, a slight variant of the above algorithm is used, where the trial energy is updated more frequently. While equilibrating, there is tendency for the number of walkers to increase exponentially as the excited eigenstates in the trial wavefunction are propagated out. The trial energy is replaced by a reference energy updated after every move of all configurations:

\begin{displaymath}
E_{\mathrm{REF}}=E_{T}-\frac{C_{\mathrm{EREF}}}{\tau} \mathrm{ln}
\left(\frac{N_w}{N_c}\right)\;\;\;,
\end{displaymath} (3.46)

where $C_{\mathrm{EREF}}$ is a system dependent parameter set to control the fluctuations in the current number of configurations $N_{w}$ compared with the target number of configurations $N_{c}$.

Several improvements to the above algorithm have been proposed. [29] These modify the Green's function where the short time approximation is expected to be least accurate (adjacent to atomic cores and near nodes), and also take account of the modified timestep due to the rejection of some of the single electron moves. These improvements give a significant reduction in the timestep error in all electron calculations. In pseudopotential calculations the improvements are much smaller. Small efficiency improvements are possible by assigning weights to each walker, and only branching or deleting walkers when the weights extend beyond a defined range. Provided sufficiently small timesteps are used with sufficiently accurate wavefunctions, these different algorithms give the same ground state energy.


next up previous contents
Next: 3.5 Summary Up: 3. Quantum Monte Carlo Previous: 3.3 Variational Monte Carlo   Contents
© Paul Kent