CONQUEST: a local orbital, large-scale DFT code
CONQUEST is a local orbital density functional theory (DFT) code, capable of massively parallel operation with excellent scaling. It uses a local orbital basis to represent the Kohn-Sham eigenstates or the density matrix. CONQUEST can be applied to atoms, molecules, liquids and solids, but is particularly efficient for large systems. The code can find the ground state using exact diagonalisation of the Hamiltonian or via a linear scaling approach. The code has demonstrated scaling to over 2,000,000 atoms and 200,000 cores when using linear scaling, and over 3,400 atoms and 850 cores with exact diagonalisation. CONQUEST can perform structural relaxation (including unit cell optimisation) and molecular dynamics (in NVE, NVT and NPT ensembles with a variety of thermostats).
Getting Started
Overview: Why CONQUEST?
There are already many DFT codes which are available under open-source licences. Here we give reasons why you might choose to use CONQUEST.
Large-scale simulations
CONQUEST is designed to scale to large systems, either using exact diagonalisation (with the multisite support function approach, we have demonstrated calculations on over 3,000 atoms) or with linear scaling (where calculations on over 2,000,000 atoms have been demonstrated). Moreover, the same code and basis sets can be used to model systems from 1 atom to more than 1,000,000 atoms.
Efficient parallelisation
CONQUEST is an inherently parallel code, with scaling to more than 800 cores demonstrated for exact diagonalisation, and nearly 200,000 cores with linear scaling. This scaling enables efficient use of HPC facilities. CONQUEST (in linear scaling mode, as well as to a certain extent for exact diagonalisation) scales best with weak scaling: fixing the number of atoms per core (or thread) and choosing a number of cores based on the number of atoms.
CONQUEST also offers some OpenMP parallelisation in linear scaling mode, with relatively low numbers of MPI threads per node, and further parallelisation performed with OpenMP.
Linear scaling DFT
The ideas of linear scaling have been current for more than twenty years, but it has proven challenging to make efficient, accurate codes to implement these ideas. CONQUEST has demonstrated effective linear scaling (with excellent parallel scaling), though is still somewhat restricted in the basis sets that can be used. For calculations beyond 5,000-10,000 atoms with DFT, linear scaling is the only option.
Basis sets
CONQUEST expresses the Kohn-Sham eigenstates or the density matrix (which are equivalent) in terms of local orbitals called support functions. These support functions are made from one of two basis sets: pseudo-atomic orbitals (PAOs) or blip functions (B-splines); the main basis functions in use in CONQUEST are the PAOs. A PAO generation code is included with the CONQUEST distribution, with well-defined and reliable default basis sets for most elements.
The simplest choice is to use one PAO for each support function (typically this allows calculations up to 1,000 atoms). For diagonalisation beyond this system size, a composite basis is used, where PAOs from several are combined into a smaller set of support functions (multi-site support functions, or MSSF). With MSSF, calculations on 3,000+ atoms are possible on HPC platforms. For linear scaling, more care is required with basis sets (more details can be found here).
Frequently Asked Questions
When should I use CONQUEST?
You can use CONQUEST for any DFT simulations that you need to perform. It is efficient for small problems, though may not be as efficient as other codes (e.g. plane wave codes) because it has been designed for massively parallel operation, which brings some overhead. If you need to perform DFT calculations on large systems (several hundred atoms or beyond) or want to perform highly parallel calculations, you should definitely consider CONQUEST.
CONQUEST uses Hamann optimised norm-conserving Vanderbilt (ONCV) pseudopotentials, which can also be used by PWSCF and Abinit which allows direct comparisons between the codes.
When should I use linear scaling?
You should use linear scaling if you need to model systems with more than about 5,000 atoms, though gains are often found for smaller systems (from 1,000 atoms upwards).
Linear scaling calculations offer the prospect of scaling to significantly larger systems than traditional DFT calculations; however, they make approximations and require some care and characterisation. In particular, instead of solving for eigenvalues and eigenstates, linear scaling methods solve for the density matrix, so that energy-resolved information (e.g. DOS and band energies) are not available. To enable linear scaling, a range is also imposed on the density matrix and it is important to test the effect of this range.
Will you implement a specific feature for me?
We cannot guarantee to implement specific features, though we are always happy to take suggestions. We also welcome new developers: if there is something that you would like to see in the code, please do talk to us about joining the development effort.
How do I report a bug?
Please use the GitHub issues page. Include details of the compiler and libraries used, the version of CONQUEST, and the input and output files (if possible). We will do our best to check the bug and fix it, but cannot guarantee to help on any timescale.
How do I get help?
The Conquest mailing list (details) is the best place to get help. However, the developers cannot guarantee to answer any questions, though they will try. Bug reports should be made through the GitHub issues page.
Quick Overview
Setting up a calculation
CONQUEST requires three types of file for a calculation:
A coordinate file
Ion files (pseudopotentials)
The input file (
Conquest_input
)
Coordinates
CONQUEST works with orthorhombic unit cells (i.e. with angles between lattice vectors at ninety degrees). The coordinate file is laid out simply: lattice vectors, number of atoms, atom coordinates (along with species and movement flags). Either fractional or Cartesian coordinates can be read (the default is fractional; Cartesian coordinates require a flag to be set in the input file). CONQUEST also reads and writes PDB format coordinate files for biomolecular simulations. More information can be found in Coordinates.
Ion files
The ion files contain the pseudopotentials and pseudo-atomic orbitals
for the elements, and follow a format similar to the ion files from Siesta
(CONQUEST can read Siesta ion files). A set of default inputs to
generate ion files is available in the directory pseudo-and-pao
.
These contain pseudopotentials based on the PseudoDojo library, and
allow ion files to be produced with the basis set generation code that
is included with CONQUEST in the tools/BasisGeneration
directory.
Full details are found here.
Conquest_input
The Conquest_input
file contains all of the input flags to
control a CONQUEST run. At a minimum, the file must specify: the run
type (e.g. static
or md
); the coordinate file name; and the number
of species and the ion file names. For a well characterised
calculation, further options must be given (for instance setting
details for the calculation of the density matrix). Simple examples
are given in Example calculations and full documentation of all options can
be found in Input tags.
Go to top
Output from a calculation
The main output from CONQUEST is in a single file, named
Conquest_out
by default (this can be changed, and output can be
written to stdout
rather than a file). This file
contains details of the calculation, energies, forces and stresses and
the various electronic structure and atomic movement calculations
performed. The most important files that are produced during a run are:
Conquest_out
The output file
Conquest_warnings
A list of any warnings issued by the code (also inConquest_out
)
coord_next.dat
The updated set of atomic positions
conquest.bib
References suggested for the calculation performed
input.log
A log of input options (both set by user and defaults)
Other files are produced by different run types, and are discussed elsewhere.
Installation
You will need to download and compile the code before you can use it; we do not supply binaries.
Downloading
CONQUEST is accessed from the GitHub repository; it can be cloned:
git clone https://github.com/OrderN/CONQUEST-release destination-directory
where destination-directory
should be set by the user.
Alternatively, it can be downloaded from GitHub as a zip file and
unpacked:
https://github.com/OrderN/CONQUEST-release/archive/master.zip
Go to top
Compiling
Once you have the distribution, you will need to compile the main
Conquest code (found in the src/
directory), along with the ion file
generation code (found in the tools/
directory). Conquest requires
a working MPI installation including a Fortran90 compiler (often
mpif90
but this can vary), along with a few standard libraries:
BLAS and LAPACK (normally provided by the system vendor)
FFTW 3.x (more detail can be found at http://www.fftw.org/)
ScaLAPACK (often provided as part of an HPC system; the source code can be obtained from the netlib repository if you need to compile it)
Additionally, Conquest can use LibXC if it is available (v4.x or later).
The library locations are set in the system.make
file in the src/system
directory, along with other parameters needed for compilation. The default file
name is system.make
but you can select another file with make SYSTEM=label
which would then use the file system.label.make
in the src/system
directory.
system.<systemname>.make
files are provided for some HPC systems used by the community, but if you want to run
locally or on a different system, you will need to create an appropriate system.make
file. Use src/system/system.example.make
as a starting point.
FC
(typicallyFC=mpif90
will be all that is required)COMPFLAGS
(set these to specify compiler options such as optimisation)BLAS
(specify the BLAS and LAPACK libraries)SCALAPACK
(specify the ScaLAPACK library)FFT_LIB
(must be left as FFTW)XC_LIBRARY
(chooseXC_LIBRARY=CQ
for the internal Conquest library, otherwiseXC_LIBRARY=LibXC_v4
for LibXC v4.x, orXC_LIBRARY=LibXC_v5
for LibXC v5.x and v6.x)Two further options need to be set for LibXC:
XC_LIB
(specify the XC libraries)XC_COMPFLAGS
(specify the location of the LibXC include and module files, e.g.-I/usr/local/include
)
Once these are set, you should make the executable using make
.
The ion file generation code is compiled using the same options required for the main code.
Go to top
Multi-threading
CONQUEST can use OpenMP for multi-threading; some multi-threading is available throughout the code, while there are specific matrix multiplication routines which can use multi-threading for the linear scaling solver. The number of threads is set via the environment variable OMP_NUM_THREADS
.
Compiler flags to enable OpenMP are dependent on the vendor, but should be specified via OMPFLAGS
in the system.make
file. If compiling with OpenMP then you should also change the variable OMP_DUMMY
in the same file to be blank to enable the number of threads to be included in the output.
On some systems, the default stack size for OpenMP is set to be rather small, and this can cause a segmentation fault when running with multiple threads. We recommend testing the effect of the environment variable OMP_STACKSIZE
(and suggest setting it to 50M or larger as a first test).
Go to top
Installing with Spack
CONQUEST and all of its dependencies can be installed with Spack. The CONQUEST package requires Spack v0.21 or later. If Spack isn’t available or up to date on your system, it is relatively straightforward to install it with user permissions following the install instructions. When setting up Spack on a new system, it is recommended to configure it to use available system compilers and system packages. Once spack is installed and set up, install CONQUEST with:
spack install conquest
and load the Conquest
executable to PATH
with
spack load conquest
The build can be customized by adding options to the
Spack spec conquest
.
The CONQUEST package includes variants for OpenMP support and different matrix multiplication kernels; more details can be found in the Spack CONQUEST package.
Go to top
Example calculations
All example calculations here use diagonalisation and PAO basis sets (with a simple one-to-one mapping between PAOs and support functions).
Static calculation
We will perform a self-consistent electronic structure calculation on bulk silicon. The coordinate file that is needed is:
10.36 0.00 0.00
0.00 10.36 0.00
0.00 0.00 10.36
8
0.000 0.000 0.000 1 T T T
0.500 0.500 0.000 1 T T T
0.500 0.000 0.500 1 T T T
0.000 0.500 0.500 1 T T T
0.250 0.250 0.250 1 T T T
0.750 0.750 0.250 1 T T T
0.250 0.750 0.750 1 T T T
0.750 0.250 0.750 1 T T T
You should save this in an appropriate file (e.g. coords.dat
).
The inputs for the ion file can be found in pseudo-and-pao/PBE/Si
(for the PBE functional). Changing to that directory and running the
MakeIonFiles
utility (in tools
) will generate the file
SiCQ.ion
, which should be copied to the run directory, and renamed
to Si.ion
. The Conquest_input
file requires only a few simple
lines at its most basic:
AtomMove.TypeOfRun static
IO.Coordinates coords.dat
Grid.GridCutoff 50
Diag.MPMesh T
Diag.GammaCentred T
Diag.MPMeshX 2
Diag.MPMeshY 2
Diag.MPMeshZ 2
General.NumberOfSpecies 1
%block ChemicalSpeciesLabel
1 28.086 Si
%endblock
The parameters above should be relatively self-explanatory; the grid
cutoff (in Hartrees) sets the integration grid spacing, and can be
compared to the charge density grid cutoff in a plane wave code
(typically four times larger than the plane wave cutoff). The
Monkhorst-Pack k-point mesh (Diag.MPMeshX/Y/Z
) is a standard
feature of solid state codes; note that the grid can be forced to be
centred on the Gamma point.
The most important parameters set the number of species and give
details of what the species are (ChemicalSpeciesLabel
). For each
species label (in this case Si
) there should be a corresponding
file with the extension .ion
(again, in this case Si.ion
).
CONQUEST will read the necessary information from this file for
default operation, so no further parameters are required. This block
also allows the mass of the elements to be set (particularly important
for molecular dynamics runs).
The output file starts with a summary of the calculation requested, including parameters set, and gives details of papers that are relevant to the particular calculation. After brief details of the self-consistency, the total energy, forces and stresses are printed, followed by an estimate of the memory and time required. For this calculation, these should be close to the following:
Harris-Foulkes Energy : -33.792210321858057 Ha
Atom X Y Z
1 -0.0000000000 0.0000000000 0.0000000000
2 -0.0000000000 0.0000000000 0.0000000000
3 -0.0000000000 0.0000000000 -0.0000000000
4 0.0000000000 0.0000000000 0.0000000000
5 -0.0000000000 0.0000000000 -0.0000000000
6 0.0000000000 0.0000000000 0.0000000000
7 -0.0000000000 0.0000000000 0.0000000000
8 -0.0000000000 -0.0000000000 0.0000000000
Maximum force : 0.00000000(Ha/a0) on atom, component 2 3
X Y Z
Total stress: -0.01848219 -0.01848219 -0.01848219 Ha
Total pressure: 0.48902573 0.48902573 0.48902573 GPa
The output file ends with an estimate of the total memory and time used.
You might like to experiment with the grid cutoff to see how the energy converges (note that the number of grid points is proportional to the square root of the energy, while the spacing is proportional to one over this, and that the computational effort will scale with the cube of the number of grid points); as with all DFT calculations, you should ensure that you test the convergence with respect to all parameters.
Go to top.
Relaxation
Atomic Positions
We will explore structural optimisation of the methane molecule (a very simple example). The coordinates required are:
20.000 0.000 0.000
0.000 20.000 0.000
0.000 0.000 20.000
5
0.500 0.500 0.500 1 F F F
0.386 0.500 0.500 2 T F F
0.539 0.607 0.500 2 T T F
0.537 0.446 0.593 2 T T T
0.537 0.446 0.407 2 T T T
The size of the simulation cell should, of course, be tested carefully to ensure that there are no interactions between images. We have fixed the central (carbon) atom, and restricted other atoms to prevent rotations or translations during optimisation.
The Conquest_input
file changes only a little from before, as
there is no need to specify a reciprocal space mesh (it defaults to
gamma point only, which is appropriate for an isolated molecule). We
have set the force tolerance (AtomMove.MaxForceTol
) to a
reasonable level (approximately 0.026 eV/A). Note that the ion files
can be generated in the same way as before, and
that we assume that the ion files are renamed to C.ion
and H.ion
.
IO.Coordinates CH4.in
Grid.GridCutoff 50
AtomMove.TypeOfRun lbfgs
AtomMove.MaxForceTol 0.0005
General.NumberOfSpecies 2
%block ChemicalSpeciesLabel
1 12.00 C
2 1.00 H
%endblock
The progress of the optimisation can be followed by searching for the
string Geom
(using grep
or something similar). In this case,
we find:
GeomOpt - Iter: 0 MaxF: 0.04828504 E: -0.83676760E+01 dE: 0.00000000
GeomOpt - Iter: 1 MaxF: 0.03755566 E: -0.83755762E+01 dE: 0.00790024
GeomOpt - Iter: 2 MaxF: 0.02691764 E: -0.83804002E+01 dE: 0.00482404
GeomOpt - Iter: 3 MaxF: 0.00613271 E: -0.83860469E+01 dE: 0.00564664
GeomOpt - Iter: 4 MaxF: 0.00126136 E: -0.83862165E+01 dE: 0.00016958
GeomOpt - Iter: 5 MaxF: 0.00091560 E: -0.83862228E+01 dE: 0.00000629
GeomOpt - Iter: 6 MaxF: 0.00081523 E: -0.83862243E+01 dE: 0.00000154
GeomOpt - Iter: 7 MaxF: 0.00073403 E: -0.83862303E+01 dE: 0.00000603
GeomOpt - Iter: 8 MaxF: 0.00084949 E: -0.83862335E+01 dE: 0.00000316
GeomOpt - Iter: 9 MaxF: 0.00053666 E: -0.83862353E+01 dE: 0.00000177
GeomOpt - Iter: 10 MaxF: 0.00033802 E: -0.83862359E+01 dE: 0.00000177
The maximum force reduces smoothly, and the structure converges well.
By adjusting the output level (using IO.Iprint
for overall output,
or IO.Iprint_MD
for atomic movement) more information about the
structural relaxation can be produced (for instance, the force
residual and some details of the line minimisation will be printed for
IO.Iprint_MD 2
).
Go to top.
Cell Parameters
We will optimise the lattice constant of the bulk silicon cell that we studied for the static calculation. Here we need to change the type of run, and add one more line:
AtomMove.TypeOfRun cg
AtomMove.OptCell T
Adjust the simulation cell size to 10.26 Bohr radii in all three directions (to make it a little more challenging). If you run this calculation, you should find a final lattice constant of 10.372 after 3 iterations. The progress of the optimization can be followed in the same way as for structural relaxation, and gives:
GeomOpt - Iter: 0 MaxStr: 0.00011072 H: -0.33790200E+02 dH: 0.00000000
GeomOpt - Iter: 1 MaxStr: 0.00000195 H: -0.33792244E+02 dH: 0.00204424
GeomOpt - Iter: 2 MaxStr: 0.00000035 H: -0.33792244E+02 dH: -0.00000017
Go to top.
Simple Molecular Dynamics
We will perform NVE molecular dynamics for methane, CH4, as a simple example of how to do this kind of calculation. You should use the same coordinate file and ion files as you did for the structural relaxation, but change the atomic movement flags in the coordinate file to allow all atoms to move (the centre of mass is fixed during MD by default). Your coordinate file should look like this:
20.00000000000000 0.00000000000000 0.00000000000000
0.00000000000000 20.00000000000000 0.00000000000000
0.00000000000000 0.00000000000000 20.00000000000000
5
0.500 0.500 0.500 1 T T T
0.386 0.500 0.500 2 T T T
0.539 0.607 0.500 2 T T T
0.537 0.446 0.593 2 T T T
0.537 0.446 0.407 2 T T T
The input file should be:
IO.Coordinates CH4.in
AtomMove.TypeOfRun md
AtomMove.IonTemperature 300
AtomMove.NumSteps 100
General.NumberOfSpecies 2
%block ChemicalSpeciesLabel
1 12.00 C
2 1.00 H
%endblock
where the default timestep (0.5fs) is necessary for simulations
involving light atoms like hydrogen. The file md.stats
contains
details of the simulation, while the trajectory is output to
trajectory.xsf
which can be read by VMD among other programs. In
this simulation, the conserved quantity is the total energy (the sum
of ionic kinetic energy and potential energy of the system) which is
maintained to better than 0.1mHa in this instance. More importantly,
the variation in this quantity is much smaller than the variation in
the potential energy. This can be seen in the plot below.

Go to top.
Tutorials
We recommend that you work through, in order, the tutorials included
in the distribution in the tutorials/
directory
to become familiar with the modes of operation of the code.
NOTE In the initial pre-release of CONQUEST (January 2020) we have not included the tutorials; they will be added over the coming months.
Go to top.
Where next?
While the tutorials have covered the basic operations of Conquest, there are many more subtle questions and issues, which are given in the User Guide.
Go to top.
User Guide
Input and output
Input files
Conquest_input
All necessary input parameters should be specified in the Conquest_input
file,
including the names of the coordinate file and the ion files. This
file controls the run; there are many sensible default values for
input parameters, but you should ensure that you understand what
they mean. After a run, the full set of relevant input parameters
(whether specified by the user, or default, are available in the file input.log
).
The most common input tags are listed briefly here. Full documentation can be found in Input tags.
AtomMove.TypeOfRun
takesstatic
,md
,sqnm
,cg
IO.Coordinates
File nameIO.Iprint
0-3 (controls amount of output)*DM.SolutionMethod
diagonDiag.MPMesh
T/FDiag.MPMeshX
(andY
andZ
) NDiag.GammaCentred
T/F
Grid.GridCutoff
Energy in Ha (sets a grid spacing \(\delta x = \pi/\sqrt{2E}\) for cutoff E in Ha)AtomMove.NumSteps
NAtomMove.MaxForceTol
in Ha/bohrAtomMove.OptCell
T/F (optimises simulation cell size)General.NumberOfSpecies
N%block ChemicalSpeciesLabel
Block specifying element number, mass and ion file nameIO.FractionalAtomicCoords
T/FSpin.SpinPolarised
T/FSpin.FixSpin
T/FSpin.Magn
Difference between spin channel occupations
minE.SCTolerance
Fractional tolerance on magnitude of residual divided by number of electronsSC.KerkerPreCondition
T/F (for Kerker preconditioning of SCF)SC.MaxIters
N (maximum number of SCF iterations)
Go to top
Ion files
The ion files contain data on the different species being modelled: valence charge, pseudopotentials, pseudo-atomic orbitals (PAOs) etc. Full details on how the PAOs are used as basis functions for CONQUEST can be found in the manual section on basis sets. A utility for generating these files is provided with CONQUEST, but Siesta ion files can also be read. The CONQUEST utility uses the pseudopotentials generated by the ONCVPSP code (though note that to generate new files for CONQUEST, you will need a small patch).
A set of input files for all elements in the PseudoDojo library for
the LDA, PBE and PBEsol exchange-correlation functionals is provided in the
directory pseudo-and-pao
. This will allow you to generate ion
files for these elements easily.
The utility for generating ion files is called MakeIonFiles, and its
source code is found in the tools/BasisGeneration
directory. It
uses the same system.make
file as CONQUEST, and following
compilation the executable will be moved to the bin
directory.
The input file is Conquest_ion_input
, and the key parameters to be set for the
ion file generation are:
General.NumberOfSpecies
to specify number of species%block SpeciesLabels
to specify what the species areIn the species block (set with
%block XX
for species XX):Atom.PseudopotentialFile
to specify the input file for the ONCVPSP codeAtom.VKBFile
to specify the file that CONQUEST needs to read (included in the library of inputs, and generally namedXX.pot
for species XX)Atom.BasisSize
to specify the size of the basis; at present this can take the values:minimal
;small
;medium
; andlarge
.
These are all included in the default input files. Further fine-grained control can be applied to the basis functions; this will be documented after the pre-release of CONQUEST.
Go to top
Coordinates
The coordinates are specified in a separate file with relatively
simple format. The coordinates can be specified in fractional form
(default) or cartesian (set the input tag IO.FractionalAtomicCoords T
).
Distance units can be Bohr radii (default) or Angstroms (set the input tag
General.DistanceUnits
to Ang
). At present,
CONQUEST only handles orthorhombic unit cells.
The coordinate file is formatted as follows:
a 0.0 0.0
0.0 b 0.0
0.0 0.0 c
NAtoms
x y z species MoveX MoveY MoveZ
.
.
.
Note that the flags MoveX
etc take values T/F and indicate whether
atoms are free to move in x, y and z, respectively. The flag
species
is an integer, and selects based on species defined in the
atomic specification section of the
Conquest_input
file.
Go to top
Output files
Main output
By default, CONQUEST writes output to the Conquest_out
file
(though the filename can be set with the parameter IO.OutputFile
,
and the flag IO.WriteOutToFile
(T/F) selects output to file or
stdout
). This file contains all details of the calculation,
including energies, forces and information on the different stages of
the calculation. The output verbosity is controlled by the
IO.Iprint
family of parameters, which allows different levels of
output detail to be set for different areas of the code. For
production runs, we expect IO.Iprint 0
to give sufficient detail;
IO.Iprint 3
provides a level of detail that would normally only be
needed for debugging.
Warnings from the calculation (including indications that the
convergence should be improved, and technical issues) are written to the
Conquest_warnings
file, which should be checked after each run. The
warnings are also written to the output file at certain IO.Iprint
levels.
Go to top
Electronic structure
Different electronic structure outputs are available; in each case, the key output flag is given. Further output flags are described in Input tags.
Charge density
Band-resolved charge density (
IO.outputWF
)Density of states (
IO.writeDOS
)Atom-projected density of states (
IO.write_proj_DOS
)Atomic charges, using the Mulliken approach (
IO.AtomChargeOutput
)
The Kohn-Sham eigenvalues are output in the eigenvalues.dat
file.
The charge densities need post-processing to convert from the
standard output format to a file compatible with visualisation
(current supported formats include Gaussian CUBE file and OpenDX
files).
Note that Becke charges can be calculated if the following parameters are set:
SC.BeckeWeights T
SC.BeckeAtomicRadii T
IO.Iprint_SC 3
This method of output will be refined soon.
Go to top
Atomic structure
During structural relaxation and molecular dynamics, the atomic
structure at the end of each step is saved in the output file
coord_next.dat
. This is in the same format as the input.
Go to top
Molecular dynamics
A molecular dynamics run will generate a number of additional plain text output files:
md.stats
— summarises thermodynamic quantities at each steps
md.frames
— contains the complete physical state of the system (lattice parameters, atomic positions, velocities, forces, stress).
md.checkpoint
— data required for MD restart, namely atomic velocities and extended system variables.
md.positions
— Atomic coordinates saved at the moment of checkpointing
trajectory.xsf
— atomic coordinates save in .xsf format, which can be visualised using (for example) VMD, ifAtomMove.WriteXSF
is true..
Full details are available in Molecular Dynamics.
Go to top
Finding the ground state
Finding the electronic ground state is the heart of any DFT code. In CONQUEST, we need to consider several linked stages: the density matrix (found using diagonalisation or linear scaling); self-consistency between charge and potential; and the support functions (though these are not always optimised).
The basis functions in CONQUEST are support functions (localised functions centred on the atoms), written as \(\phi_{i\alpha}(\textbf{r})\) where \(i\) indexes an atom and \(\alpha\) a support function on the atom. The support functions are used as basis functions for the density matrix and the Kohn-Sham eigenstates:
where \(n\) is an eigenstate index and \(\mathbf{k}\) is a point in the Brillouin zone (see here for more on this). The total energy can be written in terms of the density matrix, as:
for the Hamiltonian matrix \(H\) in the basis of support functions, with the last two terms the standard Harris-Foulkes [G1, G2] correction terms.
For diagonalisation, the density matrix is made from the coefficients of the Kohn-Sham eigenstates, \(c^{n\mathbf{k}}_{i\alpha}\), while for linear scaling it is found directly during the variational optimisation of the energy.
The question of whether to find the density matrix via diagonalisation or linear scaling is a complex one, depending on the system size, the accuracy required and the computational resources available. The simplest approach is to test diagonalisation before linear scaling.
Diagonalisation
Exact diagonalisation in CONQUEST uses the ScaLAPACK library which scales reasonably well in parallel, but becomes less efficient with large numbers of processes. The computational time will scale as \(N^3\) with the number of atoms \(N\), but will probably be more efficient than linear scaling for systems up to a few thousand atoms. (Going beyond a thousand atoms with diagonalisation is likely to require the multi-site support function technique.)
To choose diagonalisation, the following flag should be set:
DM.SolutionMethod diagon
It is also essential to test relevant parameters, as described below: the k-point grid in reciprocal space (to sample the Brillouin zone efficiently); the occupation smearing approach; and the parallelisation of k-points.
Go to top
Brillouin zone sampling
We need to specify a set of discrete points in reciprocal space to approximate integrals over the Brillouin zone. The simplest approach is to use the Monkhorst-Pack approach [G3], where a grid of points is specified in all directions:
Diag.MPMesh T Diag.MPMeshX 2 Diag.MPMeshY 2 Diag.MPMeshZ 2
This grid can be forced to be centred on the gamma point (often an
important point) using the parameter Diag.GammaCentred T
.
The origin of the Monkhorst-Pack grid may also be offset by an
arbitrary vector from the origin of the Brillouin zone, by specifying:
Diag.MPShiftX 0.0 Diag.MPShiftY 0.0 Diag.MPShiftZ 0.0
Alternatively, the points in reciprocal space can be specified explicitly by giving a number of points and their locations and weights:
Diag.NumKpts 1 %block Diag.Kpoints 0.00 0.00 0.00 1.00 %endblock Diag.Kpoints
where there must be as many lines in the block as there are k-points. It is important to note that CONQUEST does not consider space group symmetry when integrating over the Brillouin zone.
Go to top.
K-point parallelization
It is possible to parallelise over k-points: to split the processes
into sub-groups, each of which is responsible for a sub-set of the
k-points. This can be very efficient, and is specified by the
parameter Diag.KProcGroups N
, where it is important that the number
of processes is an integer multiple of the number of groups N
. It
will be most efficient when the number of k-points is an integer
multiple of the number of groups.
Go to top.
Electronic occupation smearing
The occupation numbers of the eigenstates are slightly smeared near
the Fermi level, following common practice. The default smearing type
is Fermi-Dirac smearing with a temperature (in Hartrees) set with the
flag Diag.kT
which defaults to 0.001Ha.
The Methfessel-Paxton approach [G4] to occupations allows much higher smearing temperatures with minimal effect on the free energy (and hence accuracy) of the energy. This generally gives a similar accuracy with fewer k-points, and is selected as:
Diag.SmearingType 1 Diag.MPOrder 0
where Diag.MPOrder
specifies the order of the Methfessel-Paxton
expansion. It is recommended to start with the lowest order and
increase gradually, testing the effects.
Go to top.
Padding Hamiltonian matrix by setting block size
With the default setting, the size of Hamiltonian and overlap matrices is determined by the total number of support functions. It can be a prime number and timing of diagonalisation can be very slow in such cases, since the division of the matrix into small pieces is difficult.
By padding, we can change the size of Hamiltonian matrix to improve the efficiency of the diagonalisation. To set an appropriate value for the block size of the matrix, specify the following two variables.
Diag.BlockSizeR 20 Diag.BlockSizeC 20
Note that these two numbers should be the same when padding (and when using ELPA which will be introduced to CONQUEST soon). We suggest that an appropriate value is between 20 and 200, but this should be tested.
The option for padding was introduced after v1.2, and if you would like to remove it, set the following variable.
Diag.PaddingHmatrix F
Go to top.
Linear Scaling
A linear scaling calculation is selected by setting
DM.SolutionMethod ordern
. There are two essential parameters that must be
set: the range of the density matrix, and the tolerance on the
optimisation.
DM.L_range 16.0 minE.Ltolerance 1.0e-6
The tolerance is applied to the residual (the RMS value of the
gradient of the energy with respect to the density matrix). The
maximum number of iterations in the density matrix optimisation can
be set with DM.LVariations
(default 50).
At present, CONQUEST can only operate efficiently in linear scaling mode with a restricted number of support functions (though this is an area of active development). PAO basis sets of SZ and SZP size (minimal and small in the ion file generator) will run without restrictions. For larger PAO basis sets, the OSSF approach must be used, and is effective. With a blip basis there are no restrictions, though efficient optimisation is still under active development.
It is
almost always more efficient to update the charge density while
optimising the density matrix, avoiding the need for a separate
self-consistency loop. This is set by choosing
minE.MixedLSelfConsistent T
.
An essential part of a linear scaling calculation is finding the approximate, sparse inverse of the overlap matrix. Normally this will happen automatically, but it may require some tests. The key parameters are the range for the inverse (see the Atomic Specification block, and specifically the Atomic Specification block) and the tolerance applied to the inversion.
Atom.InvSRange R DM.InvSTolerance R
A tolerance of up to 0.2 can give convergence without significantly affecting the accuracy. The range should be similar to the radius of the support functions, though increasing it by one or two bohr can improve the inversion in most cases.
The input tags are mainly found in the Density Matrix section of the Input tags page.
Go to top.
Self-consistency
The normal mode of operation for CONQUEST involves an iterative search for self-consistency between the potential and the charge density. However, it is also possible to run in a non-self-consistent manner, either with a converged charge density for electronic structure analysis, or for dynamics, which will be considerably more efficient than a self-consistent calculation, but less accurate.
Self consistency is set via the following parameters:
minE.SelfConsistent T minE.SCTolerance 1E-7 SC.MaxIters 50
The tolerance is applied to the RMS value of the residual, \(R(\mathbf{r}) = \rho^{out}(\mathbf{r}) - \rho^{in}(\mathbf{r})\), integrated over all space:
where \(\mathbf{r}_l\) is a grid point and \(\Omega\) is the
grid point volume (integrals are performed
on a grid explained in Integration Grid). The maximum number
of self-consistency cycles is set with SC.MaxIters
, defaulting
to 50.
For non-self-consistent calculations, the main flag should be set as
minE.SelfConsistent F
. The charge density at each step will
either be read from a file (if the flag General.LoadRho T
is set),
or constructed from a superposition of
atomic densities. The Harris-Foulkes functional will be used to
find the energy.
Go to top.
Restarting SCF
The SCF cycle can be restarted from a previous density matrix or
charge density, which may significantly speed up convergence.
The density matrix is automatically written out in the files Kmatrix2.*
or
Lmatrix2.*
(depending on whether diagonalisation or linear scaling
is being used). These files are read in, and the initial
charge density made from them by setting the flags:
General.LoadDM T SC.MakeInitialChargeFromK T
The charge density is not written out by default; this can be changed by
setting IO.DumpChargeDensity T
which results in the files chden.nnn
being created. To read these in as the initial charge density, the flag
General.LoadRho T
should be set.
Go to top.
Advanced options
Instabilities during self-consistency are a well-known issue in electronic structure calculations. CONQUEST performs charge mixing using the Pulay approach, where the new charge density is prepared by combining the charge densities from a number of previous iterations. In general, we write:
where \(R_{i}\) is the residual at iteration \(i\), defined above. The
fraction of the output charge density that is included is governed by
the variable \(A\), which is set by the parameter
SC.LinearMixingFactor
(default 0.5). If there is instability
during the self consistency, reducing \(A\) can help (though will likely
make convergence a little slower).
It is also advisable to apply Kerker preconditioning to the residual when the system is large in any dimension. This removes long wavelength components of the residual, reducing charge sloshing. This is controlled with the following parameters:
SC.KerkerPreCondition T SC.KerkerFactor 0.1
where the Kerker factor gives the wavevector at which preconditioning starts to reduce. The Kerker preconditioning is applied to the Fourier transform of the residual, \(\tilde{R}\) as:
where \(q^2_0\) is the square of the Kerker factor and \(q\) is a wavevector. You should test values of \(q_0\) around \(\pi/a\) where \(a\) is the longest dimension of the simulation cell (or some important length scale in your system).
Go to top.
Support functions
Support functions in CONQUEST represent the density matrix, and can be simple (pseudo-atomic orbitals, or PAOs) or compound, made from simple functions (either PAOs or blips). If they are compound, made from other functions, then the search for the ground state involves the construction of this representation. Full details of how the support functions are built and represented can be found in the manual section on basis sets.
Go to top
Charged systems
CONQUEST uses periodic boundary conditions, which require overall
charge neutrality. However, charged systems can be modelled:
if an excess of electrons is specified by the user, a uniform
positive background charge is added automatically to restore overall
neutrality. At present, there are no correction schemes implemented,
so it is important to test the convergence of the energy with unit
cell size and shape. Electrons are added by setting the parameter
General.NetCharge
.
General.NetCharge 1.0
This gives the number of extra electrons to be added to the unit cell, beyond the valence electrons.
Go to top.
Spin polarisation
CONQUEST performs collinear spin calculations only. A spin-polarised
calculation is performed by setting the parameter
Spin.SpinPolarised
to T.
Users need to specify either the total initial number of spin-up and spin-down electrons in
the simulation cell (using the parameters Spin.NeUP
and
Spin.NeDN
), or the difference between the number of spin-up and
spin-down electrons (using the parameter Spin.Magn
).
The number of electrons for each spin channel can be fixed during SCF
calculations by setting the parameter Spin.FixSpin
to T (default is F).
It is possible to specify the spin occupation in the atomic charge
densities (i.e. the number of spin-up and spin-down electrons used to
build the density). This is done in the Atomic Specification
part of the Conquest_input
file. Within the atom block for
each species, the numbers of electrons should be set with
Atom.SpinNeUp
and Atom.SpinNeDn
. Note that these numbers
must sum to the number of valence electrons for the atom.
Go to top.
Examples: FM and AFM iron
A two atom ferromagnetic iron simulation might be set up using the parameters below. Note that the net spin here is S=1 \(\mu_B\) (i.e. two more electrons in the up channel than in the down), and that the net spin is not constrained.
# example of ferro bcc Fe
Spin.SpinPolarised T
Spin.FixSpin F
Spin.NeUP 9.0 # initial numbers of up- and down-spin electrons,
Spin.NeDN 7.0 # which will be optimised by a SCF calculation when Spin.FixSpin=F
%block ChemicalSpeciesLabel
1 55.845 Fe
%endblock ChemicalSpeciesLabel
An equivalent anti-ferromagnetic calculation could be set up as follows (though note that the initial specification of spin for the atoms does not guarantee convergence to an AFM ground state). By defining two species we can create spin-up and spin-down atoms (note that both species will require their own, appropriately labelled, ion file).
# example of anti-ferro bcc Fe
Spin.SpinPolarised T
Spin.FixSpin F
Spin.NeUP 8.0 # initial numbers of up- and down-spin electrons in an unit cell
Spin.NeDN 8.0 # are set to be the same
%block ChemicalSpeciesLabel
1 55.845 Fe1
2 55.845 Fe2
%endblock ChemicalSpeciesLabel
%block Fe1 # up-spin Fe
Atom.SpinNeUp 5.00
Atom.SpinNeDn 3.00
%endblock Fe1
%block Fe2 # down-spin Fe
Atom.SpinNeUp 3.00
Atom.SpinNeDn 5.00
%endblock Fe2
When using multi-site or on-site support functions in spin-polarised
calculations, the support functions can be made spin-dependent
(different coefficients for each spin channel) or not by setting
Basis.SpinDependentSF
(T/F, default is T).
Go to top.
J. Harris. Simplified method for calculating the energy of weakly interacting fragments. Phys. Rev. B, 31:1770, 1985. doi:10.1103/PhysRevB.31.1770.
W. Foulkes and R. Haydock. Tight-binding models and density-functional theory. Phys. Rev. B, 39:12520, 1989. doi:10.1103/PhysRevB.39.12520.
H. J. Monkhorst and J. D. Pack. Special points for brillouin-zone integrations. Phys. Rev. B, 13:5188, 1976. doi:10.1103/PhysRevB.13.5188.
M. Methfessel and A. T. Paxton. High-precision sampling for brillouin-zone integration in metals. Phys. Rev. B, 40:3616–3621, Aug 1989. URL: https://link.aps.org/doi/10.1103/PhysRevB.40.3616, doi:10.1103/PhysRevB.40.3616.
Go to top.
Converging Parameters
There are various important parameters in CONQUEST that affect the convergence of the total energy, and need to be tested. Integrals are calculated on a grid; the density matrix is found approximately; a self-consistent charge density is calculated; and support functions are, in some modes of operations, optimised. These parameters are described here.
Integration Grid
While many integrals are calculated analytically or on fine grids that move with the atoms, there are still some integrals that must be found numerically, and CONQUEST uses an orthorhombic, uniform grid to evaluate these integrals (this grid is also used for the Fourier transforms involved in finding the Hartree potential). The spacing of the grid will affect the accuracy of the calculation, and it is important to test the convergence of the total energy with the grid spacing.
The grid spacing can be set intuitively using an energy (which corresponds to the kinetic energy of the shortest wavelength wave that can be represented on the grid). In atomic units, \(E = k^2/2\) with \(k = \pi/\delta\) for grid spacing \(\delta\). The cutoff is set with the parameter:
Grid.GridCutoff E
where E
is an energy in Hartrees. The grid spacing can also be
set manually, by specifying the number of grid points in each
direction:
Grid.PointsAlongX N Grid.PointsAlongY N Grid.PointsAlongZ N
If setting the grid in this manner, it is important to understand a little more about the internal workings of CONQUEST. The grid is divided up into blocks (the default size is 4 by 4 by 4), and the number of grid points in any direction must correspond to an integer multiple of the block size in that direction. The block size can be set by the user:
Grid.InBlockX N Grid.InBlockY N Grid.InBlockZ N
Note that the blocks play a role in parallelisation and memory use, so that large blocks may require larger memory per process; we recommend block sizes no larger than 8 grid points in each direction. There is also, at present, a restriction on the total number of grid points in anuy direction, that it must have prime factors of only 2, 3 and 5. This will be removed in a future release.
Go to top.
Finding the density matrix
As discussed in the section on finding the ground state, the density matrix is found either with exact diagonaliation, or the linear scaling approach. These two methods require different convergence tests, and are described separately.
Diagonalisation: Brillouin Zone Sampling
The sampling of the Brillouin zone must be tested for convergence, and the parameters are described here. The convergence of charge density will be faster than detailed electronic structure such as density of states (DOS), and it will be more accurate for these types of calculations to generate a converged charge density, and then run non self-consistently (see the section on self consistency) with appropriate k-point sampling.
Go to top.
Linear Scaling
The range applied to the density matrix (DM.L_range
) determines
the accuracy of the calculation, as well as the computational time
required (as the number of non-zero elements will increase based on a
sphere with the radius of the range, the time will increase roughly
proportional to the cube of the range). In almost all circumstances,
it is best to operate with a range which converges energy
differences and forces, rather than the absolute energy. Testing
for this convergence is an essential part of the preparation for
production calculations.
The tolerance applied to the density matrix optimisation
(minE.Ltolerance
) must be
chosen to give adequate convergence of the energy and forces. The
tolerance is applied to the residual in the calculation, defined as:
The dot product uses the inverse of the overlap matrix as the metric.
The approximate, sparse inversion of the overlap matrix is performed
before the optimisation of the density matrix. The method used,
Hotelling’s method (a version of a Newton-Raphson approach) is
iterative and terminates when the characteristic quantity
\(\Omega\) increases. On termination, if \(\Omega\) is below
the tolerance DM.InvSTolerance
then the inverse is accepted;
otherwise it is set to the identity (the density matrix optimisation
will proceed in this case, but is likely to be inefficient). We
define:
where \(T\) is the approximate inverse. The range for the inverse
must be chosen (Atom.InvSRange
in the species block); by default
it is same as the support function range
(which is then doubled to give the matrix range) but can be
increased. The behaviour of the inversion with range is not simple,
and must be carefully characterised if necessary.
Go to top.
Self-consistency
The standard self-consistency approach uses the Pulay RMM method, and should be robust in most cases. It can be monitored via the residual, which is currently defined as the standard RMS difference in charge density:
where \(\rho^{in}\) is the input charge density for an iteration,
and \(\rho^{out}\) is the resulting output charge density. The
SCF cycle is terminated when this residual is less than the parameter
minE.SCTolerance
. The maximum number of iterations is set with
SC.MaxIters
(defaults to 50).
There are various further approaches and parameters which can be used
if the SCF cycle is proving hard to converge. As is standard, the
input for a given iteration is made by combining the charge density
from a certain number of previous steps (SC.MaxPulay
, default 5).
The balance between input and output charge densities from these
previous steps is set with SC.LinearMixingFactor
(default 0.5;
N.B. for spin polarised calculations,
SC.LinearMixingFactor_SpinDown
can be set separately). Reducing
this quantity may well improve stability, but slow down the rate of
convergence.
Kerker-style preconditioning (damping long wavelength charge
variations) can be selected using SC.KerkerPreCondition T
(this is
most useful in metallic and small gap systems). The preconditioning
is a weighting applied in reciprocal space:
where \(q_0\) is set with SC.KerkerFactor
(default 0.1).
This is often very helpful with slow convergence or instability.
Go to top.
Support Functions
The parameters relevant to support functions depend on the basis set that is used. In the case of pseudo-atomic orbitals (PAOs), when support functions are primitive PAOs, the only relevant parameter is the basis set size, which is set when the ion files are generated. It is important to test the accuracy of a given basis set carefully for the problem that is to be modelled.
When using multi-site support functions (MSSF), the key parameter is
the radius of the MSSF (Atom.MultisiteRange
in
the atomic specification block).
As this is increased, the accuracy of the
calculation will also increase, but with increased computational
effort. Full details of the MSSF (and related OSSF) approach are
given in the section on multi-site support functions.
For the blip basis functions, the spacing of the grid where the blips
are defined is key (Atom.SupportGridSpacing
in
the atomic specification block),
and is directly related to an equivalent plane
wave cutoff (via \(k_{bg} = \pi/\delta\) and \(E_{PW} =
k_{bg}^2/2\), where \(\delta\) is the grid spacing in Bohr radii
and \(E_{PW}\) is in Hartrees). For a particular grid spacing,
the energy will converge monotonically with support function radius
(Atom.SupportFunctionRange
in
the atomic specification block).
A small support function radius will introduce some approximation to
the result, but improve computational performance. It is vital to
characterise both blip grid spacing and support function radius in any
calculation. A full discussion of the blip function basis is found
here.
Go to top.
Basis sets
As we have mentioned in finding the ground state, the density matrix is represented by support functions. These, in turn, are made up of basis functions, and the choice of basis set and how it represents the support functions affects both the accuracy and performance of a calculation.
There are two kinds of basis functions which are used in CONQUEST to
represent the support functions: pseudo-atomic orbitals (keyword
PAOs
) which are the default; and b-splines (keyword blips
)
which allow systematic convergence at the expense of greater
complexity.
The basis set is selected as follows:
Basis.BasisSet PAOs
When the basis set is taken to be PAOs, there are three different ways to construct the support functions, discussed below:
Each support function is represented by a single PAO (primitive PAOs)
Multi-site support functions, built from PAOs on several atoms
on-site support functions, built from PAOs on one atom
Primitive PAOs are efficient for small systems. When using large PAO basis sets for systems containing more than several hundred atoms, multi-site support functions and on-site support functions will be more efficient than primitive PAOs.
Go to top.
Pseudo-atomic orbitals (PAOs)
PAOs are solutions of the Schrodinger equation for isolated atoms, using pseudopotentials, with some confinement applied. They consist of radial functions multiplied by spherical harmonics. For the valence orbitals, the radial functions are referred to as zeta (\(\zeta\)), while for unoccupied orbitals, they are termed polarisation.
A minimal PAO basis set is single-\(\zeta\) (SZ), with one radial function for each angular momentum quantum number in the valence electrons. While the cost is significantly lower than for other basis sets, the accuracy will be rather low.
The accuracy of a calculation can be improved by adding polarisation functions and multiple radial functions for different angular momentum values, though systematic improvement is rather difficult to achieve (this is straightforward with a blip function basis). The PAO utility included with CONQUEST generates basis sets with differing sizes and accuracies; full details of the performance of these basis sets can be found elsewhere [B1].
minimal (single zeta, SZ)
small (single zeta and polarisation, SZP)
medium (double zeta, single polarisation, DZP)
large (triple zeta, double polarisation, TZDP)
Go to top.
Primitive PAOs as support functions
The easiest way to prepare support functions is to use primitive PAOs
as the support functions without any modifications. In this case, the
input parameters related to the support functions are automatically
set by obtaining the information from the PAO files (.ion
files) so long as they are generated by the CONQUEST MakeIonFiles
utility, version 1.0.3 or later. No further input parameters
need to be set in Conquest_input
.
Go to top.
Multi-site support functions
Since the computational cost of Conquest scales cubically with the
number of support functions, contracting the PAOs into a smaller set
of support functions is an efficient way to
reduce the computational cost when we use large multiple-\(\zeta\)
PAO basis sets. Multi-site support functions (MSSFs)
[B2, B3] are constructed for each
atom by taking linear combinations of the atom’s
PAOs and the PAOs from neighbouring atoms within a certain range
(set with the parameter Atom.MultisiteRange
in
the atom specification block).
Multi-site support functions can be selected by setting the following parameters:
Basis.BasisSet PAOs
Basis.MultisiteSF T
Various other parameters need to be set in the
atom specification block.
The number of support functions for the atoms must be set, and is
normally equivalent to a minimal (single zeta) basis; it is set with
Atom.NumberOfSupports
.
(To use a number of support functions larger than this minimal number, the
parameter Multisite.nonminimal
needs to be set to T
.)
The range for the multi-site support functions (the PAOs of any atom within this
distance of the atom will be included in the support functions)
is set with Atom.MultisiteRange
. The accuracy of the MSSF will
improve as this range is increased, though the computational cost will
also increase; careful tests must be made to find an appropriate
range. For a minimal number of MSSF, the range must be large enough
to include other atoms, though this restriction can be removed (see
on-site support functions for more details).
As well as setting the range for the MSSFs, we need to specify an approach for finding the expansion coefficients. A reasonable set of MSSF coefficients can be found using the local filter diagonalization (LFD) method. For improved accuracy, this should be followed by variational numerical optimisation.
Go to top.
Local filter diagonalization (LFD)
In this method, which is selected by setting Multisite.LFD T
, the
MSSF coefficients are found by diagonalising the
Hamiltonian in the primitive PAO basis, for a small cluster of atoms
surrounding the target atom. The MSSF coefficients \(C\) are determined by
projecting the sub-space molecular orbitals \(C_{sub}\) around each
atom onto localized trial vectors \(t\),
\(C = C_{sub} f(\varepsilon_{sub}) C_{sub}^T S_{sub} t\)
The cluster for diagonalisation must be at least as large as the MSSF
range, but larger clusters tend to give better MSSF coefficients (at the
expense of an increased computational cost).
The LFD sub-space region is determined for each atom by setting
Atom.LFDRange
.
An example set of parameters for an MSSF calculation for bulk Si would be:
Basis.BasisSet PAOs
Basis.MultisiteSF T
Multisite.LFD T
%block ChemicalSpeciesLabel
1 28.07 Si
%endblock
%block Si
Atom.NumberOfSupports 4
Atom.MultisiteRange 8.0
Atom.LFDRange 8.0
%endblock
When calculating binding energy curves or optimising cells, a change of
lattice constant can suddenly bring a new set of atoms within the
range of the support functions. In this case, a smearing can be
applied at the edges of the range, by setting Multisite.Smear T
.
Further details are given below.
Some form of self-consistency between the MSSF and the charge density
is required (as the MSSF will determine the Hamiltonian and hence the
output charge density). At present, this is performed as a complete
SCF cycle for each set of MSSF coefficients (though this is likely to
be updated soon for improved efficiency). This is selected by default
(but can be turned off by setting the parameter Multisite.LFD.NonSCF T
).
This iterative process is not variational, but is terminated when the
absolute energy change between iterations is less than
Multisite.LFD.Min.ThreshE
, or the residual (defined in
self-consistency) is less than
Multisite.LFD.Min.ThreshD
.
An example input block for this process would be as follows:
Multisite.LFD T
Multisite.LFD.Min.ThreshE 1.0e-6
Multisite.LFD.Min.ThreshD 1.0e-6
Go to top.
Numerical optimisation
The MSSF coefficients can also be optimised by minimizing the
DFT energy with respect to the coefficients, in a variational
process. The threshold and the
maximum iteration number of the numerical optimisation are specified
by minE.EnergyTolerance
and minE.SupportVariations
. The
optimisation is based on the conjugate gradient (CG) method, and the
initial CG step size can be specified by minE.InitStep_paomin
(default is 5.0).
minE.VaryBasis T
minE.EnergyTolerance 1.0e-6
minE.SupportVariations 30
The numerical optimisation provides more accurate coefficients than
the LFD method but is usually more time consuming. Therefore, it is
generally better to start from good initial values, for example, the
coefficients calculated by LFD. When both Multisite.LFD
and minE.VaryBasis
are selected,
the initial coefficients will be calculated by LFD
and the coefficients will then be optimised.
Basis.MultisiteSF T
Multisite.LFD T
minE.VaryBasis T
If good initial coefficient values have been found in a previous
calculation, reading these from files (the base name of these files is
SFcoeffmatrix2
) and performing only the
numerical optimisation is also a good choice.
Basis.LoadCoeffs T
Basis.MultisiteSF T
Multisite.LFD F
minE.VaryBasis T
Go to top.
Advanced MSSF concepts
Smearing the edge of the support functions Here, we are concerned with changes of lattice constant which may bring new atoms inside the support function range.
We can set the smearing-function type
Multisite.Smear.FunctionType
(default=1:Fermi-Dirac, 2=Error
function), the center position of the function
Multisite.Smear.Center
(default is equal to the range of the
support functions), offset of the center position
Multisite.Smear.Shift
and the width of the Fermi-Dirac function
Multisite.Smear.Width
(default=0.1).
Selecting states from the sub-space Here, we consider how to create the MSSF themselves from the results of the sub-space diagonalisation.
The Fermi function \(f\) with \(\varepsilon_{sub}\)
Multisite.LFD.ChemP
and \(kT\) Multisite.LFD.kT
in the
equation removes the effects of the subspace molecular orbitals in
higher energy region.
In default, \(\varepsilon_{sub}\) is automatically set to the mean
value of the subspace HOMO and LUMO energies for each subspace. If
users want to modify this, set Multisite.LFD.UseChemPsub F
and the
\(\varepsilon_{sub}\) value with Multisite.LFD.ChemP
.
For the LFD trial functions \(t\), when Atom.NumberOfSupports
is equal to the number of SZ or single-zeta plus polarization (SZP),
the PAOs which have the widest radial functions for each spherical
harmonic function are chosen as the trial vectors automatically in
default.
When Atom.NumberOfSupports
is equal to the number of SZP and
Multisite.nonminimal.offset
is set, the other PAOs will have the
weight in the trial vectors with the value of
Multisite.nonminimal.offset
.
The users can also provide the trial vectors from the input file using the LFDTrialVector
block
# Trial vectors of Au (element 1) and O (element 2) atoms.
# Au: 15 PAOs (DZP) -> 6 support functions, O: 13 PAOs (DZP) -> 4 support functions.
%block LFDTrialVector
# species sf npao s s x y z d1 d2 d3 d4 d5 d1 d2 d3 d4 d5 for Au
1 1 15 0.0 1.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
1 2 15 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0 0.0
1 3 15 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0
1 4 15 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0
1 5 15 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0
1 6 15 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0
2 1 13 0.0 1.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
2 2 13 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
2 3 13 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0 0.0 0.0 0.0
2 4 13 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.0 0.0 0.0 0.0 0.0 0.0
# species sf npao s s x y z x y z d1 d2 d3 d4 d5 for O
%endblock LFDTrialVector
The first, second and third columns correspond to the indices of species, support functions for each species, and the number of PAOs for each species. The other columns provide the initial values of the trial vectors. For example, in the first line in the above example, the second s PAO is chosen as the trial vector for the first support function of Au.
Self-consistent LFD
Two further conditions are applied to end the LFD self-consistency
process. The maximum number of iterations is set with
Multisite.LFD.Min.MaxIteration
. It is also possible, as the
process is not variational, that the energy can increase as well as
decrease between iterations. If the energy increase is less than
Multisite.LFD.Min.ThreshEnergyRise
(which defaults to ten times
Multisite.LFD.Min.ThreshE
) then convergence is deemed to have been
reached.
Go to top.
On-site support functions
On-site support functions (OSSF) are similar to multi-site support
functions, but are linear combinations of PAOs only on the target atom.
In this case, Atom.MultisiteRange
should be small enough not to
include any neighboring atoms (suggested values between 0.1 to
0.5). The number of support functions must be equivalent to the number
of functions in an SZP basis (if polarisation functions are in the
basis set) or an SZ basis (if there are no polarisation functions).
The parameter Multisite.nonminimal
should be set to true if
polarisation functions are included.
The coefficients can be determined in the same was as for MSSF (with
the LFD method and/or the numerical optimisation described above). It
is likely that significant improvement in accuracy will be found with
numerical optimisation. It is also important to test the effect of
the parameter Atom.LFDRange
which should be large enough to
include several shells of neighbouring atoms.
The OSSF approach is most likely to be useful when linear scaling calculations with large basis sets are required. An example set of parmeters is found below.
Basis.BasisSet PAOs
Basis.MultisiteSF T
Multisite.LFD T
Multisite.nonminimal T
minE.VaryBasis T
# example of Si
%block Si
Atom.NumberOfSupports 9
Atom.MultisiteRange 0.1
Atom.LFDRange 8.0
%endblock
Go to top.
Blips
Blips (which are a type of piecewise continuous polynomial called a B-spline) [B4] are useful for very accurate calculations, since the basis set can be systematically improved, in the same way as a planewave basis set. However, the calculations can be expensive depending on the parameters, and the code for blip optimisation is under development. The following description, and possible keywords, may change during development.
The blips are defined on a blip grid, which is a regular cubic grid centred on the atoms, which also moves with the atoms. The basis set can be systematically improved, by increasing the support function radius and/or reducing the spacing of the blip grids. (The support grid spacing, which defines the grid for the blips, is equivalent to a plane wave cutoff; for a given support grid spacing the energy decreases variationally with support function radius.) For each species, we need to provide these two parameters, as well as the number of support functions, which should have a minimal basis size. (At present, the smallest blip-grid spacing is used for all species.)
For a given atom, we would set:
%block atom
Atom.NumberOfSupports 4
Atom.SupportFunctionRange 6.0
Atom.SupportGridSpacing 0.3
%endblock
For each atomic species, an ion file with a minimal (SZ) basis set is required for the charge density and to initialise the blips.
The blip-grid spacing is directly related to the cutoff energy of the wavefunctions in planewave calculations. For a given cutoff energy \(E_{\rm cutoff}\) in Hartree, the blip-grid spacing should be \(\frac{2\pi}{\sqrt{2 E_{\rm cutoff}}}\) in bohr. Note that the grid spacing of integration grids (or FFT grids for the charge density) should be half the spacing of the blip grid, or smaller.
It is essential to optimise the support functions (blip coefficients) in the case of blips. The tolerance and maximum number of iterations can be set with the following keywords:
minE.VaryBasis T
minE.EnergyTolerance 0.10E-07
minE.SupportVariations 30
It is not recommended, but if memory problems are encountered for
very accurate blip calculations, you may need to switch off the
preconditioning procedure for length-scale ill conditioning by setting
the parameter minE.PreconditionBlips F
Go to top.
Reading coefficients from files
The calculated linear-combination coefficients of the support
functions are stored in SFcoeffmatrix2
files for PAOs or
blip_coeffs
files for blips. Those files can be read by setting
Basis.LoadCoeffs T
in the subsequent calculations.
Go to top.
Basis Set Superposition Error
Basis set superposition error (BSSE) arises when the two monomer units come closer and the basis set localized on one unit can act as diffuse functions for the electrons from the other unit, and therefore could be responsible for the overestimation of the binding energy for the interacting systems. It is unlikely to affect blip basis calculations [B5].
To correct this BSSE, the Counterpoise (CP) correction method [B6] is used, where the artificial stabilization is controlled by enabling the atoms in monomer calculations to improve their basis sets by including the basis sets from other monomers (using so-called ghost atoms).
When systems A and B approach and make a new system AB, the typical interaction energy between A and B is calculated as:
\(E_{AB}^{int} = E_{AB}(AB) - E_A(A) - E_B(B).\)
where \(E_{AB}(AB)\) is the energy of system AB and \(E_{A}(A)\) and \(E_{B}(B)\) are the energies of isolated A and B. The lowerscript and parentheses correspond to the system and its structure, respectively.
Now, the estimate for the amount of artificial stabilization of A coming from the extra basis functions from B is:
\(E_{A}^{BSSE} = E_{A\bar{B}}(AB) - E_A(A\text{ in }AB),\)
where \(\bar{A}\) and \(\bar{B}\) are the ghost atoms, which have basis functions, but no potential or charge density. \(E_{A\bar{B}}(AB)\) is the energy of system A with the basis sets from ghost-atom system B in the AB structure. \(E_A(A\text{ in }AB)\) is the energy of system A in the AB structure but without system B (neither basis functions nor atoms). Therefore, the subtraction corresponds to how much system A is stabilized by the basis function of B.
Similarly, for monomer B,
\(E_{B}^{BSSE} = E_{\bar{A}B}(AB) - E_B(B\text{ in }AB),\)
Subtracting the BSSE part of A and B units from the typical interaction energy mentioned above, the counterpoise corrected interaction energy without BSSE \((E_{AB}^{int,CP})\) will be:
\(E_{AB}^{int,CP} = E_{AB}^{int} - E_{A}^{BSSE} - E_{B}^{BSSE}.\)
Practically, to calculate \(E_{A\bar{B}}(AB)\), the basis
functions of B should be placed on atomic centers of B, however with
zero nuclear charge and mass. This can be performed in CONQUEST by
specifying negative masses for the ghost atoms in B in the block
ChemicalSpeciesLabel
of the input file:
%block ChemicalSpeciesLabel
1 1.01 A
2 -1.01 B
%endblock
Go to top.
David R. Bowler, Jack S. Baker, Jack T. L. Poulton, Shereif Y. Mujahed, Jianbo Lin, Sushma Yadav, Zamaan Raza, and Tsuyoshi Miyazaki. Highly accurate local basis sets for large-scale DFT calculations in conquest. Jap. J. Appl. Phys., 58:100503, 2019. doi:10.7567/1347-4065/ab45af.
Ayako Nakata, David R. Bowler, and Tsuyoshi Miyazaki. Efficient Calculations with Multisite Local Orbitals in a Large-Scale DFT Code CONQUEST. J. Chem. Theory Comput., 10:4813, 2014. doi:10.1039/C5CP00934K.
Ayako Nakata, David Bowler, and Tsuyoshi Miyazaki. Optimized multi-site local orbitals in the large-scale DFT program CONQUEST. Phys. Chem. Chem. Phys., 17:31427, 2015. doi:10.1021/ct5004934.
E. Hernández, M. J. Gillan, and C. M. Goringe. Basis functions for linear-scaling first-principles calculations. Phys. Rev. B, 55:13485–13493, 1997. doi:10.1103/PhysRevB.55.13485.
P.D. Haynes, C.-K. Skylaris, A.A. Mostofi, and M.C. Payne. Elimination of basis set superposition error in linear-scaling density-functional calculations with local orbitals optimised in situ. Chem. Phys. Lett., 422:345 – 349, 2006. doi:10.1016/j.cplett.2006.02.086.
S.F. Boys and F. Bernardi. The calculation of small molecular interactions by the differences of separate total energies. some procedures with reduced errors. Mol. Phys., 19:553, 1970. doi:10.1080/00268977000101561.
Go to top.
Electronic Structure
CONQUEST can be used to produce a wide variety of information on the electronic structure of different systems, including: density of states (DOS) and atom-projected DOS (or pDOS); band-resolved charge density; band structure; and electronic polarisation. Many of these are produced with the post-processing code using a converged charge density. All of these (at present) require the exact diagonalisation approach to the ground state; linear scaling solutions are not possible.
Converged charge density
In most cases (except polarisation) the data required is produced by a non-self-consistent calculation which reads in a well-converged charge density. The convergence is mainly with respect to Brillouin zone sampling, but also self-consistency (a tight tolerance should be used). The basic procedure is:
Perform a well-converged calculation, writing out charge density (ensure that the Brillouin zone is well sampled, the SCF tolerance is tight (
minE.SCTolerance
) and that the flagIO.DumpChargeDensity T
is set)Perform a non-self-consistent calculation for the quantity desired (set
minE.SelfConsistent F
andGeneral.LoadRho T
to read and fix the charge density) using an appropriate Brillouin zone samplingRun the appropriate post-processing to generate the data
However, note that the charge density often converges much faster with respect to Brillouin zone sampling than the detailed electronic structure, so the use of a non-self-consistent calculation is more efficient. Often it is most efficient and accurate to use a very high density k-mesh for the final, non-SCF calculation, but a lower density k-mesh to generate the charge density (which converges faster with respect to Brillouin zone sampling than DOS and other quantities).
Go to top.
Density of states
The total density of states (DOS) is generated from the file eigenvalues.dat
which is
written by all diagonalisation calculations. See density of states for
details on parameters which can be set.
The atom-projected DOS resolves the total DOS into contributions from individual atoms
using the pseudo-atomic orbitals, and can further decompose this into l-resolved or
lm-resolved densities of states. It requires the wave-function coefficients, which will be generated
by setting IO.write_proj_DOS T
; further analysis is performed in post-processing.
Go to top.
Band structure
The band structure along a series of lines in reciprocal space can be generated. See post-processing for more details.
Go to top.
Band-resolved densities
A band-resolved density is the quantity \(\mid \psi_n(\mathbf{r}) \mid^2\)
for the \(n^{\mathrm{th}}\) Kohn-Sham eigenstate (we plot density because
the eigenstates are in general complex). It requires wavefunction coefficients
which are generated by setting IO.outputWF T
. Full details are found in
the band density section of the post-processing
part of the manual.
Go to top.
Electronic Polarisation
The electronic polarisation (the response of a material to an
external electric field) can be calculated using the approach
of Resta [ES1] by setting the tag General.CalcPol T
.
The direction in which polarisation is found is set using the tag
General.PolDir
(choosing 1-3 gives x, y or z, respectively, while
choosing 0 gives all three directions, though this is normally not
recommended).
The Resta approach is a version of the modern theory of polarisation (MTP) (perhaps better known in the method of King-Smith and Vanderbilt [ES2]) where the polarisation is found as:
where \(\mathrm{L}\) is a simulation cell length along an appropriate direction and \(V\) is the simulation cell volume. This approach is only valid in the large simulation cell limit, with \(\Gamma\) point sampling (e.g. for BaTiO3, a minimum of 3x3x3 formula units is needed, though this is perhaps a little too small).
As with all calculations in the MTP, the only valid physical quantity is a change of polarisation between two configurations. A very common quantity to calculate is the Born effective charge (BEC), which is defined as \(Z^{*}_{k,\alpha\beta} = V\partial P_{\alpha}/\partial u_{k,\beta}\) for species \(k\) and Cartesian directions \(\alpha\) and \(\beta\). It is most easily calculated by finding the change in polarisation as one atom (or one set of atoms in a sublattice) is moved a small amount.
Go to top.
R. Resta. Macroscopic polarization from electronic wave functions. Int. J. Quantum Chem., 75:599–606, 1999. doi:10.1002/(SICI)1097-461X(1999)75:4/5%3C599::AID-QUA25%3E3.0.CO;2-8.
R. D. King-Smith and D. Vanderbilt. Theory of polarization of crystalline solids. Phys. Rev. B, 47:1651–1654, 1993. doi:10.1103/PhysRevB.47.1651.
Go to top.
Structural relaxation
This section describes how to find the zero-Kelvin equilibrium atomic structure, given a starting structure with non-zero forces and/or stresses. CONQUEST can employ a variety of algorithms to minimise energy with respect to atomic positions, including: stabilised quasi-Newton method (SQNM); L-BFGS; conjugate gradients (CG); and damped molecular dynamics (both MDMin and FIRE approaches). The minimisation of energy or enthalpy with respect to cell vectors is restricted to conjugate gradients at present, though L-BFGS will be implemented.
Setting AtomMove.WriteXSF T
for all flavours of optimisation will dump the
trajectory to the file trajectory.xsf
, which can be visualised using VMD and XCrysDen.
Setting AtomMove.AppendCoords T
will append the structure at each step to UpdatedAtoms.dat
in the format of a
CONQUEST structure input.
For the SQNM, L-BFGS and conjugate gradients relaxations, the progress of the calculation can be
monitored by searching for the word GeomOpt
; grepping will print the
following:
$ grep GeomOpt Conquest_out
GeomOpt - Iter: 0 MaxF: 0.00329282 H: -0.14168571E+03 dH: 0.00000000
GeomOpt - Iter: 1 MaxF: 0.00331536 H: -0.14168995E+03 dH: 0.00424155
GeomOpt - Iter: 2 MaxF: 0.00350781 H: -0.14168997E+03 dH: 0.00001651
GeomOpt - Iter: 3 MaxF: 0.00504075 H: -0.14169161E+03 dH: 0.00164389
GeomOpt - Iter: 4 MaxF: 0.00725611 H: -0.14169172E+03 dH: 0.00010500
GeomOpt - Iter: 5 MaxF: 0.01134145 H: -0.14169329E+03 dH: 0.00157361
GeomOpt - Iter: 6 MaxF: 0.01417229 H: -0.14169385E+03 dH: 0.00056077
GeomOpt - Iter: 7 MaxF: 0.01434628 H: -0.14169575E+03 dH: 0.00190304
GeomOpt - Iter: 8 MaxF: 0.01711197 H: -0.14170001E+03 dH: 0.00425400
GeomOpt - Iter: 9 MaxF: 0.02040556 H: -0.14170382E+03 dH: 0.00381110
GeomOpt - Iter: 10 MaxF: 0.01095167 H: -0.14170752E+03 dH: 0.00370442
In this example, MaxF is the maximum single force component, H is the enthalpy and dH is the change in enthalpy.
Go to top.
Ionic relaxation
To optimise the ionic positions with respect to the DFT total energy, the following flags are essential:
AtomMove.TypeOfRun sqnm
AtomMove.MaxForceTol 5e-4
AtomMove.ReuseDM T
The parameter AtomMove.TypeOfRun
can take the values sqnm
, lbfgs
or
cg
for iterative optimisation. All three algorithms are robust and
relatively efficient in most instances; SQNM [SR1] is recommended in most cases,
though if the initial forces are large it may be worth performing quenched
MD to reduce them (see below) before applying SQNM. The
parameter AtomMove.MaxForceTol
specifies the force
convergence criterion in Ha/bohr, i.e. the calculation will terminate
when the largest force component on any atom is below this value.
The parameter
AtomMove.ReuseDM
specifies that the density matrix (the K-matrix for
diagonalisation or L-matrix for O(N) calculations) from the
previous step will be used as an initial guess for the SCF cycle after
propagating the atoms; this should generally decrease the number of SCF cycles
per ionic step. When using CG, the line minimiser can be chosen: safe
uses a robust though sometimes slow line minimiser; backtrack
uses a simple back-tracking line minimiser (starting with a step size of 1 and reducing if necessary to ensure the energy goes down); adapt
uses an adaptive back-tracking line minimiser (which increases the starting step size if the energy goes down on the first step). In many cases the back-tracking line minimiser is more efficient, though the efficiency of the adaptive approach varies with problem.
If the self-consistency tolerance is too low, the optimisation may fail to
converge with respect to the force tolerance; this may necessitate a tighter
minE.SCTolerance
for diagonalisation (also possibly
minE.LTolerance
for O(N) calculations). A grid which is too
coarse can also cause problems with structural relaxation to high tolerances.
For large initial forces or problematic cases where the relaxation algorithms fail to find a downhill search direction, it may be worth trying quenched molecular dynamics, which propagates the equations of motion following a simple NVE approach, but resets the velocities to zero when the dot product of force and velocity is zero.
AtomMove.TypeOfRun md
AtomMove.QuenchedMD T
AtomMove.MaxForceTol 5e-4
AtomMove.ReuseDM T
The FIRE algorithm [SR2] is a variant of quenched MD that has been shown to outperform conjugate gradients in some circumstances.
AtomMove.TypeOfRun md
AtomMove.FIRE T
AtomMove.MaxForceTol 5e-4
AtomMove.ReuseDM T
Go to top.
Simulation cell optimisation
The simulation cell can be optimised with respect to enthalpy with fixed fractional
coordinates (AtomMove.OptCellMethod 1
) using the following input:
AtomMove.TypeOfRun cg
AtomMove.OptCell T
AtomMove.OptCellMethod 1
AtomMove.ReuseDM T
AtomMove.EnthalpyTolerance 1E-5
AtomMove.StressTolerance 0.1
Note that stress is in GPa and enthalpy is in Ha by default.
Go to top.
Combined optimisation
For simple crystals, the fractional ionic coordinates vary trivially with
changes in the simulation cell lengths; however for more complicated systems such as
molecular crystals and amorphous materials, it is necessary simultaneously relax
the ionic positions and simulation cell lengths (recalling that CONQUEST only
allows orthorhombic unit cells). This can be done by setting
AtomMove.OptCellMethod 2
or AtomMove.OptCellMethod 3
AtomMove.TypeOfRun cg
AtomMove.OptCell T
AtomMove.OptCellMethod 2
AtomMove.ReuseDM T
AtomMove.MaxForceTol 5e-4
AtomMove.EnthalpyTolerance 1E-5
AtomMove.StressTolerance 0.1
Note that stress is in GPa and enthalpy is in Ha by default.
The enthalpy will generally converge much more rapidly than the force
and stress, and that it may be necessary to tighten minE.SCTolerance
(diagonalisation) or minE.LTolerance
(order(N)) to reach the force
and stress tolerance, if it is even possible. For combined optimisation,
we recommend using AtomMove.OptCellMethod 2
,
which uses a simple but robust double-loop minimisation: a full ionic
relaxation (using either cg or sqnm) followed by a full simulation cell
relaxation (using cg). While this may be less efficient than optimising all
degrees of freedom simultaneously, it is much more robust. It is also possible
to optimise cell vectors and atomic positions simultaneously, using AtomMove.OptCellMethod 3
,
but this should be monitored carefully, as it can be unstable.
Go to top.
Bastian Schaefer, S. Alireza Ghasemi, Shantanu Roy, and Stefan Goedecker. Stabilized quasi-newton optimization of noisy potential energy surfaces. J. Chem. Phys., 142(3):034112, 2015. doi:10.1063/1.4905665.
E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumbsch. Structural Relaxation Made Simple. Phys. Rev. Lett., 97:2897, 2006. doi:10.1103/PhysRevLett.97.170201.
Go to top.
Molecular Dynamics
CONQUEST can perform molecular dynamics both when the density matrix is computed using diagonalisation and O(N), the latter allowing dynamical simulations of (but not limited to) tens of thousands of atoms. The equations of motion are integrated using the velocity Verlet method in the case of the microcanonical ensemble (NVE), and modifications thereof for the canonical (NVT) and isobaric-isothermal (NPT) ensembles, the details of which can be found in Molecular Dynamics: Theory. In addition to converging the parameters for the electronic structure calculations, the following points must also be considered.
Go to top.
Self-consistency tolerance and XL-BOMD
The convergence of the electronic structure is important in MD, as
insufficient convergence can be responsible for “drift” in the
conserved quantity of the dynamics. Although the molecular dynamics
integrators used in CONQUEST are time reversible, the SCF procedure
is not. Therefore tight convergence (minE.SCTolerance
for
diagonalisation, minE.LTolerance
for linear scaling) is
necessary. In the case of diagonalisation, SCF tolerance of 1E-6
is
typically enough to negate the drift. However, extended-Lagrangian
Born-Oppenheimer MD (XL-BOMD) [MD1], currently only
implemented for O(N), essentially makes the SCF component of the MD
time-reversible by adding the electronic degrees of freedom to the
Lagrangian, relaxing the constraint on minE.LTolerance
—
although it is still somewhat dependent on the ensemble. In the NVE
and NVT ensembles, a L-tolerance of 1E-5
has been found to be
sufficient to give good energy conservations, decreasing to 1E-6
in the NPT ensemble. The following flags are required for XL-BOMD:
DM.SolutionMethod ordern
AtomMove.ExtendedLagrangian T
Go to top.
Restarting
Assuming the calculation ended gracefully, it can easily be restarted by setting,
AtomMove.RestartRun T
This will do several things: it will read the atomic coordinates from
md.position
and read the md.checkpoint
file, which contains the
velocities and extended system (Nose-Hoover chain and cell) variables. Depending
on the value of DM.SolutionMethod
, it will read the K-matrix files
(diagon
) or the L-matrix files (ordern
), and if XL-BOMD is being used,
the X-matrix files. Finally, it will append new data to the md.stats
and
md.frames
files, but it will overwrite all other files, including
Conquest_out
. Note that this flag is equivalent to setting the following:
General.LoadL T
SC.MakeInitialChargeFromK T
XL.LoadL T
In addition to the files mentioned above, CONQUEST will try to read the K-matrix
from Kmatrix2.i00.*
when using diagonalisation or the L-matrix from
Lmatrix2.i00.*
when using O(N), and Xmatrix2.i0*.*
if the
extended-Lagrangian formalism is used. Note that metadata for these files is
stored in InfoGlobal.i00.dat
which is also required when restarting. If the
calculation ended by hitting the walltime limit, the writing of these matrix
files may have been interrupted, rendering them unusable. In this case, the
calculation can be restarted by setting the above flags to F
after setting
AtomMove.RestartRun T
. Setting the flag General.MaxTime
to some number
of seconds less (say 30 minutes) than the calculation wall time limit will force
the calculation to stop gracefully, preventing the aforementioned situation.
Go to top.
Visualising the trajectory
Setting the flag AtomMove.WriteXSF T
dumps the coordinates to the file
trajectory.xsf
every AtomMove.OutputFreq
steps. The .xsf file can be
read using VMD. A small VMD script,
view.vmd
is included with the code, and can be invoked using,
vmd -e view.vmd
assuming the vmd executable is in your path.
Go to top.
TDEP output
CONQUEST molecular dynamics data can be used to perform lattice dyanmical
calculations using the Temperature Dependent Effective Potential (TDEP) code. Setting the flag MD.TDEP
T
will make conquest dump configurations, forces and metadata in a format
readable by TDEP.
Go to top.
Non-Hamiltonian dynamics
Canonical (NVT) ensemble
The thermostat is set using the MD.Thermostat
flag, and can take the values
svr
(stochastic velocity rescaling) and nhc
(Nose-Hoover
chain). These thermostats generate the correct canonical ensemble
phase space distribution, and both give a conserved quantity that
allows the quality of the dynamics to be monitored.
Stochastic velocity rescaling
AtomMove.IonTemperature 300.0
MD.Ensemble nvt
MD.Thermostat svr
MD.tauT 10
While the NHC uses chaotic sensitivity to initial conditions to achieve better
ergodicity, the SVR thermostat [MD2] uses a judiciously chosen stochastic force
coupled to a weak scaling thermostat to correctly generate the
canonical phase space distribution. The MD.tauT
parameter gives
the coupling timescale; the velocity scaling factor is modified by a
factor \(\Delta t/\tau\), so a larger \(\tau\) results in a
more slowly varying temperature. While some characterisation of the
system is recommended, values of \(\tau\) around 20–200fs are
reasonable. To reproduce a simulation, the random number
generator seed can be set with the General.RNGSeed <integer>
flag.
Nose-Hoover chain
AtomMove.IonTemperature 300.0
MD.Ensemble nvt
MD.Thermostat nhc
MD.nNHC 5
MD.nYoshida 5
MD.tauT 30
When thermostatting using a Nose-Hoover chain [MD3, MD4, MD5], it may be necessary to set a
couple more flags. MD.nNHC
sets the number of thermostats in the chain (the
default of 5 is generally sensible), and MD.nYoshida
determines the order of
Yoshida-Suzuki integration. This is essentially a higher level integration
scheme that can improve energy conservation in cases when rapid changes in the
Nose-Hoover thermostat velocity is causing integration errors. Note that
MD.tauT
means something different to the SVR case. A good guess is
the time period of the highest frequency motion of the system in fs; however, in
the NVT ensemble, the energy conservation is not very sensitive to this value.
The NHC masses can also be set manually using the following block.
MD.CalculateXLMass F
MD.nNHC 5
%block MD.NHCmass
5 1 1 1 1
%endblock
Go to top.
Isobaric-Isothermal (NPT) ensemble
There is one implemented barostat at present, the extended system, Parrinello-Rahman [MD6]. At present the barostat should be treated as a beta-version implementation, which will be fully characterised and made robust for the full release of the code.
Parrinello-Rahman
AtomMove.IonTemperature 300.0
AtomMove.TargetPressure 10.0
MD.Ensemble npt
MD.Thermostat nhc
MD.Barostat pr
MD.nNHC 5
MD.nYoshida 5
MD.tauT 100
MD.tauP 200
MD.PDrag 10.0
The Parrinello-Rahman barostat generates the correct ensemble, but can
be subject to low frequency “ringing” fluctuations in the
temperature and pressure that can destabilise the system or slow equilibration.
Unlike in the NVT ensemble, this combination of barostat and thermostat is
very sensitive to the choice of both MD.tauT
and MD.tauP
; note that
their values are somewhat higher in this case, since integration errors in the
NHC tend to be more severe due to coupling of the cell and atomic motions. They
are dependent on the system, so it is advised that you find a combination of
these parameters that gives the best energy conservation. The cell is
thermostatted using a separate Nose-Hoover chain to the atoms by default, but
they can be controlled with the same chain by setting MD.CellNHC F
. An ad
hoc drag factor specified by MD.PDrag
reduces the thermostat and cell
velocities at every timestep to damp out the ringing fluctuations. In this case,
they are reduced by \(10/200 \simeq 5\%\), which strictly speaking breaks the NPT
dynamics, but not significantly, and the stability is significantly improved.
Note that the NPT ensemble can also be generated correctly by thermostatting
using the SVR thermostat, although the meaning of the parameter MD.tauT
is
different in this case, as in NVT dynamics.
Postprocessing tools
Details of Python post-processing tools for CONQUEST can be found in Molecular dynamics analysis.
Go to top.
A. M. N. Niklasson. Extended Born-Oppenheimer Molecular Dynamics. Phys. Rev. Lett., 100:123004, 2008. doi:10.1103/PhysRevLett.100.123004.
G. Bussi, D. Donadio, and M. Parrinello. Canonical sampling through velocity rescaling. J. Chem. Phys., 126:014101, 2007. doi:10.1063/1.2408420.
S. Nosé. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys., 81:511, 1984. doi:10.1063/1.447334.
W. G. Hoover. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A, 31:1695, 1985. doi:10.1103/PhysRevA.31.1695.
G. J. Martyna, M. L. Klein, and M. Tuckerman. Nosé–hoover chains: the canonical ensemble via continuous dynamics. J. Chem. Phys., 97:2635, 1992. doi:10.1063/1.463940.
M. Parrinello and A. Rahman. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys., 52:7182–7190, December 1981. doi:10.1063/1.328693.
Go to top.
Post-processing CONQUEST output
Introduction
The utility PostProcessCQ
allows users to post-process the output
of a CONQUEST calculation, to produce structure files, densities of
states, and charge density, band
densities and STM images as CUBE
files (which can be read by the freely available VESTA code).
There are a number of different analyses which can be performed:
coordinate conversion (to formats which can be plotted); conversion of
total charge density to CUBE file format; production of band-resolved
(optionally k-point resolved) densities in CUBE file format; simple
Tersoff-Hamann STM simulation; and calculation of densities of states,
including projected DOS. You should ensure that all the files
produced during the CONQUEST run are available for the post-processing
(including eigenvalues.dat
, chden.NNN
, make_blk.dat
or
hilbert_make_blk.dat
and ProcessNNNNNNNWF.dat
and
ProcessSijNNNNNNNWF.dat
as applicable) as well as the input files.
Note that the utility reads the Conquest_input
file, taking some
flags from the CONQUEST run that generated the output, and some
utility-specific flags that are detailed below.
Note also that projected DOS, band density and STM simulation are not at present compatible with multi-site support functions (MSSF), though we hope to implement this soon.
Go to top.
Coordinate conversion
Set Process.Job coo
to output a coordinate file for further
processing or plotting. The utility will read the file specified by
Process.Coordinates
(which defaults to the file specified by
IO.Coordinates
). The output format is selected by specifying the
Process.CoordFormat
tag. The default output format is XYZ (which
adds a .xyz
suffix to the file name) using xyz
. The CASTEP
.cell
output format can also be selected using cell
. We plan to
expand this conversion to other formats in the future.
Note that for a structural relaxation or molecular dynamics
calculation, if you do not specify Process.Coordinates
then the
IO.Coordinates
file, which will be converted, will be the input
structure, not the output structure. Parameters that can be set are:
Process.Coordinates string (default: IO.Coordinates value)
Process.CoordFormat string (default: xyz; options: xyz, cell)
Go to top.
Charge density
Setting Process.Job
to cha
, chg
or den
will convert
the files chden.NNN
which are written by CONQUEST to a cube file.
The processing will use the files chden.NNN
, Conquest_input
and hilbert_make_blk.dat
or raster_make_blk.dat
. Parameters
that can be set include:
Process.ChargeStub string (default: chden)
The ChargeStub simply defines the filename which will be read, and used for output.
Note that to output the chden.NNN
files from CONQUEST, you must
set the flag IO.DumpChargeDensity T
in the CONQUEST run.
Go to top.
Band density
Setting Process.Job
to ban
produces band densities from wave
function coefficients output by CONQUEST. The CONQUEST run must have
the following tags set:
IO.outputWF T
A set of bands whose coefficients are output are specified either with an energy range (the default is to produce all bands):
IO.WFRangeRelative T/F
IO.min_wf_E real (Ha)
IO.max_wf_E real (Ha)
or with a list of bands:
IO.maxnoWF n
%block WaveFunctionsOut
n entries, each a band number
%endblock
The wavefunction range can be relative to the Fermi level
(IO.WFRangeRelative T
) otherwise it is absolute. Either of these
will produce a file containing all eigenvalues at all k-points
(eigenvalues.dat
) and a series of files containing the
wavefunction expansion coefficients for the selected bands
(ProcessNNNNNNNWF.dat
). These files are output as binary
(unformatted) by default (this can be changed by setting
IO.MatrixFile.BinaryFormat F
before the CONQUEST run) and will be
read using the same format (it is important to check this!).
From these wavefunction coefficient files, band densities can be produced in post-processing, using similar tags; either a range:
Process.min_wf_E real (Ha)
Process.max_wf_E real (Ha)
Process.WFRangeRelative T/F
or an explicit list of bands:
Process.noWF n
%block WaveFunctionsProcess
n entries, each a band number
%endblock
Note that the bands to be processed must be a subset of the bands
output by CONQUEST. The bands can be output summed over k-points, or
at individual k-points, by setting Process.outputWF_by_kpoint
to
F
or T
respectively.
Go to top.
Tersoff-Hamann STM simulation
Setting Process.Job ter
will use a very simple Tersoff-Hamann
approach to STM simulation, summing over band densities between the
Fermi level and the bias voltage (this is often surprisingly
accurate). The following parameters can be set:
STM.BiasVoltage real (eV)
STM.FermiOffset real (eV)
Process.MinZ real (Bohr)
Process.MaxZ real (Bohr)
Process.RootFile string (default: STM)
The FermiOffset
tag allows the user to shift the Fermi level (to simulate
charging or an external field). The height of the simulation cell
in which the STM image is calculated is set by the MinZ
and
MaxZ
tags, and the filename by the RootFile
tag.
Go to top.
Density of states (DOS)
Setting Process.Job dos
will produce a total density of states
(DOS) for the system, using the eigenvalues output by CONQUEST. The
following parameters can be set:
Process.min_DOS_E real (Ha, default lowest eigenvalue)
Process.max_DOS_E real (Ha, default highest eigenvalue)
Process.sigma_DOS real (Ha, default 0.001)
Process.n_DOS integer (default 1001)
The limits for the DOS are set by the first two parameters (note that
CONQUEST will output all eigenvalues, so the limits on these are set
by the eigenspectrum). The broadening applied to each state is set by
sigma_DOS
, while the number of bins is set by n_DOS
. The
integrated DOS is also calculated; the user can choose whether this
is the total integrated DOS (i.e. from the lowest eigenvalue,
regardless of the lower limit for DOS) or just the local integrated
DOS (i.e. over the interval specified for the DOS) by setting
Process.TotalIntegratedDOS
to T
or F
, respectively.
We recommend that, for accurate DOS, CONQUEST should be run
non-self-consistently with a very high k-point density, after reading
in a well-converged input charge density: set minE.SelfConsistent
F
and General.LoadRho T
(which will require that the converged
charge density is written out by CONQUEST by setting IO.DumpChargeDensity T
).
Go to top.
Atom-projected DOS
Setting Process.Job pdos
will produce a total density of states as
above, as well as the density of states projected onto the individual
atoms. Given support functions \(\phi_{i\alpha}(\mathbf{r})\)
which are the basis functions of the Kohn-Sham eigenstates
\(\psi_{n}(\mathbf{r}) = \sum_{i\alpha}
c^{n}_{i\alpha}\phi_{i\alpha}(\mathbf{r})\), then the projection of a
given state, \(n\), onto an atom \(i\) can be written as
\(\sum_{\alpha j\beta} c^{n}_{i\alpha}
S_{i\alpha,j\beta}c^{n\mathbf{k}}_{j\beta}\). The projected DOS is
constructed using these projections.
If using pseudo-atomic orbitals (PAOs) as the basis set, then the atom-projected DOS can be further resolved by angular momentum (either just \(l\) or both \(l\) and \(m\)). If using pseudo-atomic orbitals (PAOs) with multi-site support functions or blip functions then it is not possible to decompose the DOS any further (in future, it may be possible to resolve the MSSF coefficients into the individual PAOs, and hence decompose pDOS by angular momentum). To output the necessary coefficients to produce atom-projected DOS, a CONQUEST run must be performed with the following parameters set:
IO.writeDOS T
IO.write_proj_DOS T
As for the DOS, very high Brillouin zone sampling is required for
accurate projected DOS, which is most efficiently generated using a
converged charge density and a non-self-consistent calculation with
much higher k-point density. CONQUEST will produce the wavefunction
files (ProcessNNNNNNNWF.dat
and ProcessSijNNNNNNNWF.dat
) as
binary (unformatted) by default (change using the flag
IO.MatrixFile.BinaryFormat F
).
Once the files have been generated by CONQUEST, the output can be processed by setting the output tag:
Process.Job pdos
This is all that is needed for the simplest output. The number of bins and smearing of the peaks can be set using:
Process.sigma_DOS 0.002
Process.n_DOS 10001
To resolve the DOS by angular momentum as well as by atom, then the following flags can be set:
Process.pDOS_l_resolved T
Process.pDOS_lm_resolved T
Note that only one of these is needed, depending on what level of resolution is required. At present, angular momentum resolution is only available for the PAO basis set (not MSSF or blips) though it is under development for the MSSF basis (by projection onto the underlying PAO basis).
The energy range for the projected DOS can also be specified:
Process.min_DOS_E -0.35
Process.max_DOS_E 0.35
Process.WFRangeRelative T
where the final tag sets the minimum and maximum values relative to the Fermi level.
If you only want to produce pDOS for a few atoms, then you can set
the variable Process.n_atoms_pDOS
and list the atoms you want
in the block pDOS_atoms
:
Process.n_atoms_pDOS 2
%block pDOS_atoms
1
12
%endblock
Go to top.
Band structure
The band structure of a material can be generated by CONQUEST by performing
a non-self-consistent calculation, after reading a well-converged charge density:
set minE.SelfConsistent F
and General.LoadRho T
(remember that to write
a converged charge density from CONQUEST you set IO.DumpChargeDensity T
).
The k-points required can be specified as lines of points in k-space;
setting Diag.KspaceLines T
enables this (replacing the usual MP mesh), while the number of lines
(e.g. Gamma to L; L to X; would be two lines) is set with Diag.NumKptLines
and the number of points along a line with Diag.NumKpts
. The k-point lines
themselves are set with a block labelled Diag.KpointLines
which should have
two entries (starting and finishing k-points) for each k-point line.
(In constructing the k-point list, CONQUEST will automatically remove any duplicate
points, so that the output can be plotted smoothly.) So to create
a bandstructure from X-\(\Gamma\)-L-X (3 lines: X-\(\Gamma\); \(\Gamma\)-L;
L-X) with 11 points in each line, you would use the following input:
Diag.KspaceLines T
Diag.NumKptLines 3
Diag.NumKpts 11
%block Diag.KpointLines
0.5 0.0 0.0
0.0 0.0 0.0
0.0 0.0 0.0
0.5 0.5 0.5
0.5 0.5 0.5
0.5 0.0 0.0
%endblock
After running CONQUEST, setting Process.Job bst
and running the post-processing
will read the resulting
eigenvalues.dat
file, and produce a file BandStructure.dat
. The x-axis will
be the k-point index by default, but specifying Process.BandStrucAxis
(taking
value n
for index, x
, y
or z
for a single direction in k-space,
or a
to give all k-point coordinates) will allow you to control this. Limits
on the energies to select the bands produced can be set
with Process.min_DOS_E
and Process.max_DOS_E
.
Go to top.
Managing Conquest with ASE
Below we give an introduction how to setup the ASE environment with respect to CONQUEST repository along with a few examples of ASE/Conquest capabilities. We assume that a python script or jupyter-notebook is used.
Setup
Environment variables
The script will need to set environmental variables specifying the
locations of the CONQUEST executable Conquest
, and if required, the basis
set generation executable MakeIonFiles
and pseudopotential database.
These variables are:
ASE_CONQUEST_COMMAND
: the Conquest executable command including MPI/openMPI prefix.CQ_PP_PATH
: the PAO path directory to where are located the the.ion
files.(optional)
CQ_GEN_BASIS_CMD
: the PAO generation executableMakeIonFiles
.
Given the Conquest root directory CQ_ROOT
, initialisation might look to something like
import os
CQ_ROOT = 'PATH_TO_CONQUEST_DIRECTORY'
os.environ['ASE_CONQUEST_COMMAND'] = 'mpirun -np 4 '+CQ_ROOT+'/bin/Conquest'
os.environ["CQ_GEN_BASIS_CMD"] = CQ_ROOT+'/bin/MakeIonFiles"
os.environ['CQ_PP_PATH'] = CQ_ROOT+'/pseudo-and-pao/'
Go to top.
Pseudopotential/PAO files
Conquest atomic pseudotential and basis functions are store in the .ion
files which will ne referred as to PAO files. Provided the pseudopotential files .pot
available in CQ_PP_PATH
,
automatic generation of numerical PAOs is possible using the program MakeIonFiles
available from the Conquest package.
Provided the PAO files, the basis set is specified through a python dictionary, for example:
basis = {'O' : {'file': 'O_SZP.ion'},
'H' : {'file': 'H_SZP.ion'},
'C' : {'file': 'C_SZP.ion'}}
In this case they are all assumed to be obtained from Hamann pseudopotentials,
which are the default. Knowing the the exchange and correlation functional <XC>
from the Conquest input (vide infra) and the chemical symbol <X>, the Calculator
will search the .ion
file in different places:
CQ_PP_PATH
CQ_PP_PATH/lib/
CQ_PP_PATH/<XC>/<X>/
including the current directory and the ASE working directory (vide infra). If your PAO file is located in a different place you can include the path in the basis dictionary:
basis = {'O' : {'file': 'O_SZP.ion',
'directory': '<PATH_TO_FILE>'},
'H' : {'file' : 'H_SZP.ion'},
'C' : {'file' : 'C_SZP.ion'}}
For generating the PAO files, the keyword gen_basis
should be set to True
(default is False
) and the size be provided (default is medium
).
For instance:
basis = {'O' : {'gen_basis' : True,
'basis_size': 'small'},
'C' : {'gen_basis' : True,
'basis_size': 'medium'},
'H' : {'gen_basis' : True,
'basis_size': 'large'}}
will create the O.ion
, C.ion
and H.ion
files where small
, medium
and large
are default size basis set. You are allowed to choose the functional and
to add options for basis set generation:
basis = {'H' : {'gen_basis' : True,
'basis_size': 'small',
'xc' : 'LDA',
'Atom.Perturbative_Polarised': False}}
Note
Only Hamann pseudopotentials for LDA, PBE and PBEsol are available within the CONQUEST distribution. For using other functionals see Generating new pseudopotentials.
Warning
Generating polarised PAOs for some atoms can be problematic (mainly group I
and II). Please review carefuly the MakeIonFiles
input files
named Conquest_ion_input
which are collected in CQ_PP_PATH/<XC>/<X>/
if you are not sure about what you are doing, and check your PAOs.
Go to top.
CONQUEST Calculator
The CONQUEST Calculator class can be invoked from the ase Calculator set as described in the example below:
from ase.calculators.conquest import Conquest
A minimal example is given below for setting the CONQUEST Calculator (named calc
)
of the ASE Atoms object
named struct:
from ase.calculators.conquest import Conquest
from ase.build import bulk
struct = bulk('NaCl', crystalstructure='rocksalt', a=5.71, cubic=True)
basis = {'Cl' : {'file' : 'Cl.ion'}, 'Na' : {'file' : 'Na.ion'}}
calc = Conquest(basis=basis,atoms=struct)
or, equivalently,
from ase.calculators.conquest import Conquest
from ase.build import bulk
struct = bulk('NaCl', crystalstructure='rocksalt', a=5.71, cubic=True)
basis = {'Cl' : {'file' : 'Cl.ion'}, 'Na' : {'file' : 'Na.ion'}}
struct.calc = Conquest(basis=basis)
In basic calculate mode (compute energy
), the Calculator comes with 3 methods:
write_input()
:this function will setup the input files. For CONQUEST, the PAO basis will be generated/copied with respect to the dictionary key/value pairs, and
Conquest_input
file including the calculation parameters will be written, a long with thecoordinate
file, containing the lattice vectors (in Bohr Unit) and atomic positions (in fractional coordinates).
execute()
:this function execute the calculation. For CONQUEST, it will launch the
ASE_CONQUEST_COMMAND
setup in the environment varaibles.
read_results()
:this function post-process the the output file. For CONQUEST, the
energy
,forces
,stress
andeigenvalues
will be extracted from theConquest_out_ase
output file.
Note
The funtion read_results()
operate on the Conquest_out_ase
file.
This output file is not created by default by CONQUEST. If you want to post-process
a calculation with an input generated by hand you must add IO.WriteOutToASEFile True
in conquest_input
.
The indirect way for managing CONQUEST calculation with ASE is:
struct.calc = Conquest(basis=basis)
struct.calc.write_input(struct)
struct.calc.execute()
struct.calc.read_results(struct)
where struct.calc.execute()
can be ignored when, for instance, the calculation
is performed on a supercomputer and the output file is then copied back to the current
directory for post-processing.
The direct way is simply:
struct.calc = Conquest(basis=basis)
struct.calc.calculate(struct)
or, equivalently,
struct.calc = Conquest(basis=basis)
struct.get_potential_energy()
Go to top.
Keywords for generating the Conquest_input file
In principle all the Conquest input parameters
can be added to Conquest_out_ase
using key/value pairs in a dictionary. There are 3 class of parameters:
mandatory : they are parsed to the Calculator and have no defaults ; there are mandatory.
important : they are parsed to the Calculator they can be freely modified. Some of them are pure ASE keywords.
defaults : they are set as defaults ; some of them must not be modified. They are read by the Calculator through a dictionay
conquest_flags
.
Mandatory keywords
keyword |
type |
default value |
description |
---|---|---|---|
|
|
None |
an atoms object constructed either via ASE or read from an input |
|
|
None |
a dictionary specifying the pseudopotential/basis files |
Important keywords
keyword |
CONQUEST equivalence |
type |
default value |
description |
---|---|---|---|---|
|
None |
|
None |
directory used for storing input/output and calculation files |
|
None |
|
None |
basename for working files (only used by ASE, eg. NEB) |
|
None |
|
None |
k-points grid ; converted to CONQUEST Monkhorst-Pack grid |
|
|
|
100 |
integration grid in Ha |
|
|
|
‘PBE’ |
exchange and correlation functional |
|
|
|
True |
choose either SCF or non-SCF |
|
|
|
1e-6 |
Self-consistent-field convergence tolerance in Ha |
|
|
|
1 |
spin polarisation: 1 for unpolarized or 2 for polarised |
|
None |
|
None |
other CONQUET keyword arguments |
Defaults keywords
keyword |
type |
default value |
description |
---|---|---|---|
|
|
True |
write ASE output file ; must always be True when using ASE for post-processing |
|
|
1 |
verbose for the output ; must always be 1 when using ASE for post-processing |
|
|
‘diagon’ |
‘diagon’ stands for diagonalisation other is ‘ordern’ (base on density matrix) |
|
|
‘Hamann’ |
kind of pseudopotential other type are ‘siesta’ and ‘abinit’ |
|
|
50 |
maximum number SCF cycles |
|
|
‘static’ |
‘static’ stands for single (non)SCF other are ‘md’ or optimisation algorithms. |
|
|
1 |
1 for Methfessel-Paxton ; 0 for Fermi-Dirac |
|
|
0.001 |
smearing temperature in Ha |
Some examples
An example of more advanced Calculator setup is given below for a SCF calculation on BCC-Na where for a PBE calculation using a k-point grid of \(6\times 6\times6\) using the Fermi-Dirac distribution for the occupation with a smearing of 0.005 Ha:
struct = bulk('Na', crystalstructure='bcc', a=4.17, cubic=True)
basis = {'Na' : {'file' : 'NaCQ.ion'}}
conquest_flags = {'Diag.SmearingType': 0,
'Diag.kT' : 0.005}
struct.calc = Conquest(directory = 'Na_bcc_example',
grid_cutoff = 90.0,
self_consistent= True,
xc = 'PBE',
basis = basis,
kpts = [6,6,6],
nspin = 1,
**conquest_flags)
struct.get_potential_energy()
Finally, defaults and other input flags can be defined in a new dictionary, and passed as an expanded set of keyword arguments.
conquest_flags = {'DM.SolutionMethod' : 'ordern',
'DM.L_range' : 12.0,
'minE.LTolerance' : 1.0e-6}
Here is an example, combining the above. We set up a cubic diamond cell containing 8 atoms, and perform a single point energy calculation using the order(N) method (the default is diagonalisation, so we must specify all of the order(N) flags). We don’t define a basis set, instead providing keywords that specify that a minimal basis set should be constructed using the MakeIonFiles basis generation tool.
import os
from ase.build import bulk
from ase.calculators.conquest import Conquest
CQ_ROOT = 'PATH_TO_CONQUEST_DIRECTORY'
os.environ['ASE_CONQUEST_COMMAND'] = 'mpirun -np 4 '+CQ_ROOT+'/bin/Conquest'
os.environ["CQ_GEN_BASIS_CMD"] = CQ_ROOT+'/bin/MakeIonFiles"
os.environ['CQ_PP_PATH'] = CQ_ROOT+'/pseudo-and-pao/'
diamond = bulk('C', 'diamond', a=3.6, cubic=True) # The atoms object
conquest_flags = {'DM.SolutionMethod' : 'ordern', # Conquest keywords
'DM.L_range' : 12.0,
'minE.LTolerance' : 1.0e-6}
basis = {'C': {'basis_size' : 'minimal', # Generate a minimal basis
'gen_basis' : True}
calc = Conquest(grid_cutoff = 80, # Set the calculator keywords
xc = 'PBE',
self_consistent=True,
basis = basis,
nspin = 1,
**conquest_flags)
diamond.set_calculator(calc) # attach the calculator to the atoms object
energy = diamond.get_potential_energy() # calculate the potential energy
Go to top.
Multisite support functions
Multisite support functions require a few additional keywords in the atomic species block, which can be specified as follows:
basis = {'C': {"basis_size": 'medium',
"gen_basis": True,
"pseudopotential_type": "hamann",
"Atom.NumberofSupports": 4,
"Atom.MultisiteRange": 7.0,
"Atom.LFDRange": 7.0}}
Note that we are constructing a DZP basis set (size medium) with 13 primitive
support functions using MakeIonFiles
, and contracting it to multisite basis
of 4 support functions. The calculation requires a few more input flags, which
are specified in the other_keywords
dictionary:
other_keywords = {"Basis.MultisiteSF": True,
"Multisite.LFD": True,
"Multisite.LFD.Min.ThreshE": 1.0e-7,
"Multisite.LFD.Min.ThreshD": 1.0e-7,
"Multisite.LFD.Min.MaxIteration": 150,
}
Go to top.
Loading the K/L matrix
Most calculation that involve incrementally moving atoms (molecular dynamics, geometry optimisation, equations of state, nudged elastic band etc.) can be made faster by using the K or L matrix from a previous calculation as the initial guess for a subsequent calculation in which that atoms have been moved slightly. This can be achieved by first performing a single point calculation to generate the first K/L matrix, then adding the following keywords to the calculator:
other_keywords = {"General.LoadL": True,
"SC.MakeInitialChargeFromK": True}
These keywords respectively cause the K or L matrix to be loaded from file(s)
Kmatrix.i**.p*****
, and the initial charge density to be constructed from
this matrix. In all subsequent calculations, the K or L matrix will be written
at the end of the calculation and used as the initial guess for the subsequent
ionic step.
Go to top.
Equation of state
The following code computes the equation of state of diamond by doing single
point calculations on a uniform grid of the a
lattice parameter. It then
interpolates the equation of state and uses matplotlib
to generate a plot.
import scipy as sp
from ase.build import bulk
from ase.io.trajectory import Trajectory
from ase.calculators.conquest import Conquest
# Construct a unit cell
diamond = bulk('C', 'diamond', a=3.6, cubic=True)
basis = {'C': {"basis_size": 'minimal',
"gen_basis": True}}
calc = Conquest(grid_cutoff = 50,
xc = "PBE",
basis = basis,
kpts = [4,4,4])
diamond.set_calculator(calc)
cell = diamond.get_cell()
traj = Trajectory('diamond.traj', 'w') # save all results to trajectory
for x in sp.linspace(0.95, 1.05, 5): # grid for equation of state
diamond.set_cell(cell*x, scale_atoms=True)
diamond.get_potential_energy()
traj.write(diamond)
from ase.io import read
from ase.eos import EquationOfState
configs = read('diamond.traj@0:5')
volumes = [diamond.get_volume() for diamond in configs]
energies = [diamond.get_potential_energy() for diamond in configs]
eos = EquationOfState(volumes, energies)
v0, e0, B = eos.fit()
import matplotlib
eos.plot('diamond-eos.pdf') # Plot the equation of state
Go to top.
External tools
Post-processing for charge density, band density, DOS, STM
The utility PostProcessCQ
allows users to post-process the output
of a CONQUEST calculation, to produce the charge density, band
densities, DOS and STM images in useful forms. It is described fully here.
Molecular dynamics analysis
Several scripts that may be helpful with postprocessing molecular dynamics are
included with CONQUEST. The can be found in the tools
directory, and the
executables are plot_stats.py
, md_analysis.py
and heat_flux.py
. They
have the following dependencies:
Python 3
Scipy/Numpy
Matplotlib
If Python 3 is installed the modules can be added easily using pip3 install
scipy
etc.
These scripts should be run in the calculation directory, and will automatically
parse the necessary files, namely Conquest_input
, input.log
,
md.stats
and md.frames
assuming they have the default names. They will
also read the CONQUEST input flags to determine, for example, what ensemble is
used, and process the results accordingly.
Go to top.
Plotting statistics
usage: plot_stats.py [-h] [-c] [-d DIRS [DIRS ...]]
[--description DESC [DESC ...]] [--skip NSKIP]
[--stop NSTOP] [--equil NEQUIL] [--landscape]
[--mser MSER_VAR]
Plot statistics for a CONQUEST MD trajectory
optional arguments:
-h, --help show this help message and exit
-c, --compare Compare statistics of trajectories in directories
specified by -d (default: False)
-d DIRS [DIRS ...], --dirs DIRS [DIRS ...]
Directories to compare (default: .)
--description DESC [DESC ...]
Description of graph for legend (only if using
--compare) (default: )
--skip NSKIP Number of equilibration steps to skip (default: 0)
--stop NSTOP Number of last frame in analysis (default: -1)
--equil NEQUIL Number of equilibration steps (default: 0)
--landscape Generate plot with landscape orientation (default:
False)
--mser MSER_VAR Compute MSER for the given property (default: None)
Running plot_stats.py --skip 200
in your calculation will generate a plot
which should resemble the example below, skipping the first 200 steps. This
example is a molecular dynamics simulation of 1000 atoms of bulk silicon in the
NPT ensemble, at 300 K and 0.1 GPa.

The four plots are respectively the breakdown of energy contributions, the
conserved quantity, the temperature and the pressure, the last of which is only
included for NPT molecular dynamics. Several calculations in different
directories can be compared using plot_stats.py --compare -d dir1
dir2 --description "dir1 description" "dir2 description"
. The following
example compares the effect of changing the L tolerance in the above simulation.
Note that the contents of the description field will be in the legend of the
plot.

Go to top.
MD analysis
usage: md_analysis.py [-h] [-d DIRS [DIRS ...]] [--skip NSKIP]
[--stride STRIDE] [--snap SNAP] [--stop NSTOP]
[--equil NEQUIL] [--vacf] [--msd] [--rdf] [--stress]
[--nbins NBINS] [--rdfwidth RDFWIDTH] [--rdfcut RDFCUT]
[--window WINDOW] [--fitstart FITSTART] [--dump]
Analyse a CONQUEST MD trajectory
optional arguments:
-h, --help show this help message and exit
-d DIRS [DIRS ...], --dirs DIRS [DIRS ...]
Directories to compare (default: .)
--skip NSKIP Number of equilibration steps to skip (default: 0)
--stride STRIDE Only analyse every nth step of frames file (default:
1)
--snap SNAP Analyse Frame of a single snapshot (default: -1)
--stop NSTOP Number of last frame in analysis (default: -1)
--equil NEQUIL Number of equilibration steps (default: 0)
--vacf Plot velocity autocorrelation function (default:
False)
--msd Plot mean squared deviation (default: False)
--rdf Plot radial distribution function (default: False)
--stress Plot stress (default: False)
--nbins NBINS Number of histogram bins (default: 100)
--rdfwidth RDFWIDTH RDF histogram bin width (A) (default: 0.05)
--rdfcut RDFCUT Distance cutoff for RDF in Angstrom (default: 8.0)
--window WINDOW Window for autocorrelation functions in fs (default:
1000.0)
--fitstart FITSTART Start time for curve fit (default: -1.0)
--dump Dump secondary data used to generate plots (default:
False)
The script md_analysis.py
script performs various analyses of the trajectory
by parsing the md.frames` file. So far, these include the radial distribution
function, the velocity autocorrelation function, the mean squared deviation, and
plotting the stress. For example, the command,
md_analysis.py --rdf --stride 20 --rdfcut 8.0 --nbins 100 --dump --skip 200 --stop 400
computes the radial distribution function of the simulation in the first example from every 20th time step (every 10 fs in this case), stopping after 400 steps, with a cutoff of 8.0 A, and the histogram is divided into 100 bins.

Go to top.
CONQUEST structure file analysis
usage: structure.py [-h] [-i INFILE] [--bonds] [--density] [--nbins NBINS]
[-c CUTOFF [CUTOFF ...]] [--printall]
Analyse a CONQUEST-formatted structure
optional arguments:
-h, --help show this help message and exit
-i INFILE, --infile INFILE
CONQUEST format structure file (default:
coord_next.dat)
--bonds Compute average and minimum bond lengths (default:
False)
--density Compute density (default: False)
--nbins NBINS Number of histogram bins (default: 100)
-c CUTOFF [CUTOFF ...], --cutoff CUTOFF [CUTOFF ...]
Bond length cutoff matrix (upper triangular part, in
rows (default: None)
--printall Print all bond lengths (default: False)
The script structure.py
can be used to analyse a CONQUEST-formatted
structure file. This is useful to sanity-check the bond lengths or density,
since an unphysical structure is so often the cause of a crash. For example, the
bond lengths can be computed with
structure.py --bonds -c 2.0 3.0 3.0
where the -c
flag specifies the bond cutoffs for the bonds 1-1, 1-2 and 2-2,
where 1 is species 1 as specified in Conquest_input
and 2 is species 2. The
output will look something like this:
Mean bond lengths:
O-Si: 1.6535 +/- 0.0041 (24 bonds)
Minimum bond lengths:
O-Si: 1.6493
Go to top.
Atomic Simulation Environment (ASE)
ASE is a set of
Python tools for setting up, manipulating, running, visualizing and analyzing
atomistic simulations. ASE contains a CONQUEST interface, also
called Calculator so that it can be used to calculate energies
, forces
and stresses
as inputs to other calculations such as Phonon
or NEB that
are not implemented in CONQUEST. ASE is a versatil tool to manage CONQUEST
calculations without pain either:
in a direct way where pre-processing, calculation and post-processing are managed on-the-fly by ASE,
or in an indirect way where the calculation step is performed outside the workflow, ie. on a supercomputer.
The ASE repository containing the Conquest calculator can be found here. Detailed documentation on how to manage Conquest calculations with ASE is available here.
Go to top.
Error codes
Error codes generated by different CONQUEST routines will be collated here, along with explanations of the root cause and suggested fixes. This will form part of the full release, but is not implemented in the pre-release.
Generating PAOs
Introduction
CONQUEST includes a utility for generating the PAO basis files (MakeIonFiles
with source code in the directory tools/BasisGeneration
), and we also provide pseudopotential files (from the PseudoDojo database). The input files will generate a reasonable default basis set (though we can offer no guarantees: users must test the accuracy and convergence of their basis sets). Here we will discuss how to generate both default and custom basis sets. Full details of the basis sets can be found in a recent paper [GP1].
Default basis sets
To generate basis functions, radii for the PAOs must be specified; by default, the utility will set these radii automatically. The radii can be set so that the different shells either share the same radii, or share an energy shift associated with confinement. The default behaviour is to generate basis sets where shells share radii (to change this, add the line Atom.Cutoffs energy
to the input file). There are four default basis set sizes:
minimal (single zeta, SZ)
small (single zeta and polarisation, SZP)
medium (double zeta, single polarisation, DZP)
large (triple zeta, triple polarisation, TZTP)
Generally, reasonable results will be obtained with a medium (DZP) basis, though this should always be tested. Minimal and small basis sets are much faster (and are the only basis sets compatible with linear scaling), but less reliable. The large basis set will be slower (often it is twice the size of the medium basis set, so diagonalisation will be up to eight times slower) but more reliable and accurate.
We note that Group I and II atoms are a little problematic: the standard approach for most other elements may produce a somewhat limited basis set, so we have created a more accurate, customised input file for these elements (with the exception of Na and Mg, where the pseudopotential does not include l=2 components, so the default approach is all that is possible). These should be tested carefully.
Go to top.
Specifying basis sets
The generation utility gives the user complete control over the basis sets that are produced. As an example, we reproduce below the input file for strontium (Sr) and discuss the layout.
%block Sr
Atom.PseudopotentialFile Sr.in
Atom.VKBFile Sr.pot
Atom.Perturbative_Polarised F
Atom.PAO_N_Shells 5
Atom.BasisBlock SrBlock
%endblock
%block SrBlock
# n, l, number of zetas
4 0 1
4 1 1
5 0 2
5 1 1
4 2 1
# Radii for PAOs (bohr)
4.0
5.0
10.1 5.7
10.1
10.1
%endblock
In this case, we specify five shells (combinations of n and l) via the Atom.PAO_N_Shells
tag, with polarisation functions found simply by solving the Schrodinger equation in the usual way (the alternative, perturbative polarisation, is discussed below). We must then specify a block that defines these shells (Atom.BasisBlock SrBlock
). Within that block, we give the number of zeta functions for each (n,l) pair (specified as a line n l nzeta
) followed by the radii for the zeta functions.
Setting radii for the different shells is a complex process which requires considerable time and care, with an extensive literature; we cannot provide significant help, but only make suggestions. In the first instance, the default radii are a good starting point. Note that the default setting Atom.Cutoffs radii
averages the radii between shells, while Atom.Cutoffs energy
finds different radii for each shell (so that the energy change due to confinement is the same for all shells). We recommend starting from one of these sets of radii, and then testing and optimising the radii against some key properties of the system.
A common approach to the generation of polarisation functions (i.e. unoccupied states) is to perturb a valence state (typically the highest energy valence state) to generate a function with angular momentum increased by one; this is the default behaviour. In this case, the radii for the polarisation state should be the same as the shell being polarised (so for Si, we would perturb the 3p (n=3, l=1) state to get the 3d (n=3, l=2) state), and at present the same number of zeta functions must be specified for the polarisation shell as for the unperturbed shell. For instance, for Si:
%block Si
Atom.PseudopotentialFile Si.in
Atom.VKBFile Si.pot
Atom.PAO_N_Shells 3
Atom.BasisBlock SiBlock
%endblock
%block SiBlock
# n, l, number of zetas
3 0 2
3 1 2
3 2 2
# Radii for PAOs (bohr)
8.0 4.0
8.0 4.0
8.0 4.0
%endblock
The perturbative option can be turned off by specifying Atom.Perturbative_Polarised F in the input file. (Note that in the strontium example above we have specified two polarisation shells, so cannot use the perturbative approach.)
By default, the utility calculates radii which are shared between shells; it is possible to specify instead shared energy shifts using Atom.Cutoffs energy
, but this can only be done for valence shells, and so must use the perturbative polarisation approach for polarisation functions.
Go to top.
Compiling
To compile the code, the same system.make
can be used as is specified for the main code. Once this is done, simply issue the ccommand make
in the tools/BasisGeneration
directory. The resulting executable will be placed in the bin
directory.
Go to top.
Generating new pseudopotentials
CONQUEST is supplied with a complete set of pseudopotentials for the elements in the PseudoDojo database (covering LDA, PBE and PBEsol exchange-correlation functionals). In order to generate new pseudopotential files, users will need the Hamann pseudopotential code ONCVPSP v3.3.1 (the current release) and the patch file Conquest_ONCVPSP_output.patch
which is in the tools
directory. After patching and compiling the Hamann code (to patch the code, copy the patch to the ONCVPSP src
directory, and issue the command patch -p0 < Conquest_ONCVPSP_output.patch
; we cannot provide any support for this) the oncvpsp.x
utility will generate a file VPS.dat
which should be renamed (something like element.pot
as in the CONQUEST pseudopotential files) and specified in the input file using the Atom.VKBFile
tag.
Go to top.
David R. Bowler, Jack S. Baker, Jack T. L. Poulton, Shereif Y. Mujahed, Jianbo Lin, Sushma Yadav, Zamaan Raza, and Tsuyoshi Miyazaki. Highly accurate local basis sets for large-scale DFT calculations in conquest. Jap. J. Appl. Phys., 58:100503, 2019. doi:10.7567/1347-4065/ab45af.
Go to top.
Tutorials
Introductory Tutorials
These introductory tutorials will give you an overview of how to run Conquest, the files and parameter settings required, and what output to expect.
Bulk silicon: input, output and SCF
We start with a very basic introduction to the input
required for CONQUEST, the output generated, and the self-consistency
(SCF) procedure; it uses the same system as the first of the examples
in the manual, but provides more detail. The files are found in
docs/tutorials/Introductory_1
.
CONQUEST requires the following files to run:
The input file:
Conquest_input
A coordinates file (name set in
Conquest_input
; no default)Ion files (suffix
.ion
), which provide the pseudopotentials and pseudo-atomic orbitals (PAOs)
The input file requires the user to provide a certain amount of information. The minimal file that is provided for this tutorial gives most of these:
# Input/Output
IO.Title Bulk Si 8 atoms static
IO.Coordinates ionpos.dat
# General Parameters
General.NumberOfSpecies 1
%block ChemicalSpeciesLabel
1 28.0850 Si_SZ
%endblock
# Moving Atoms
AtomMove.TypeOfRun static
# Finding the density matrix
DM.SolutionMethod diagon
# k-points
Diag.GammaCentred T
Diag.MPMesh T
Diag.MPMeshX 2
Diag.MPMeshY 2
Diag.MPMeshZ 2
The key entries are:
the coordinate file (
IO.Coordinates
);the number of species (
General.NumberOfSpecies
);the specification for the species (the block
ChemicalSpeciesLabel
gives the atomic mass and the ion file name for all species);the type of run (
AtomMove.TypeOfRun
which defaults tostatic
)
The Brillouin zone sampling must be investigated carefully, as for
all periodic electronic structure calculations. The Monkhorst-Pack
mesh (Diag.MPMesh
) offers a convenient way to do this systematically.
The job title is purely for reference. Further parameters are
discussed in the next tutorial
The coordinate file
IO.Coordinates
The number of species
General.NumberOfSpecies
The ion files for the species
The basic input file
The output
Changing the output level and destination
Controlling the SCF (tolerance and iterations, options)
Bulk silicon: parameters to converge
The files that are needed
Coordinates
Ion files
Input file:
Conquest_input
Integration grid
Brillouin zone sampling
Possibly basis set size
Bulk silicon: analysis
The files that are needed
Coordinates
Ion files
Input file:
Conquest_input
Total DOS
Atom-projected DOS
Band structure output
Charge density and bands
Atomic charges
Structural Relaxation Tutorials
We now turn to structural relaxation, covering both atomic relaxation and simulation cell relaxation.
Atomic relaxation: lbfgs
Parameters that are needed
The output
Files in
Structural_relaxation_1
Atomic relaxation: Conjugate gradients
Parameters that are needed
The output
Files in
Structural_relaxation_2
Atomic relaxation: Quenched MD
Parameters that are needed
The output
Files in
Structural_relaxation_3
Cell relaxation: Conjugate gradients
Parameters that are needed
The output
Files in
Structural_relaxation_4
Molecular Dynamics Tutorials
These tutorials introduce molecular dynamics (MD) in CONQUEST, though not the topic of MD itself, for which you should consult an appropriate textbook.
Molecular dynamics: NVE
Parameters that are needed
The output
Files in
MD_1
Molecular dynamics: NVT with SVR
Parameters that are needed
The output
Files in
MD_2
Molecular dynamics: NVT with Nose-Hoover chains
Parameters that are needed
The output
Files in
MD_3
Molecular dynamics: NPT with SVR and Parrinello-Rahman
Parameters that are needed
The output
Files in
MD_4
Molecular dynamics: NPT with Nose-Hoover chains
Parameters that are needed
The output
Files in
MD_5
Molecular dynamics: Continuing a run
Parameters that are needed
The output
Files in
MD_6
Basis Function Tutorials
We now introduce details of the basis sets used in CONQUEST, and how they are specified and optimised.
MSSF: local filter diagonalisation
Parameters that are needed
The output
Files in
Basis_sets_1
MSSF: Optimisation
Parameters that are needed
The output
Files in
Basis_sets_2
Advanced Tutorials
These tutorials cover more advanced topics in CONQUEST, and assume a reasonable familiarity and confidence with the general operation of the code.
Initialising spins
Parameters that are needed
The output
Files in
Advanced_1
An introduction to linear scaling
Parameters that are needed
The output
Files in
Advanced_2
Theory
Background on energy, forces and stress
A number of different ways of formulating the energy exist in Conquest at the moment, involving both self-consistent and non-self-consistent densities and potentials, both with and without the neutral atom potential. With self-consistency, all formulations should give the same result, though numerical issues may give small differences; without self-consistency the Harris-Foulkes functional is more accurate.
Self-consistent calculations
We define the energy in Conquest in two ways that are equivalent at the self-consistent ground state. The Harris-Foulkes energy is given as:
where the first term is the band structure energy, equivalent to the sum over the energies of the occupied states, the second two terms compensate for double counting and the final term gives the ion-ion interaction:
The Hamiltonian is defined as:
where the operators are the kinetic energy, the local and non-local pseudopotentials, the Hartree potential, defined as \(V_{Ha} = \int d\mathbf{r}^\prime n(\mathbf{r}^\prime)/\mid \mathbf{r} - \mathbf{r}^\prime\mid\), and the exchange-correlation potential. The alternative form, often known as the DFT energy, is:
with the Hartree energy defined as usual:
along with the exchange-correlation energy:
For the Harris-Foulkes and DFT energies to be equal, it is easy to see that the double counting correction terms in the Harris-Foulkes formalism must be:
and
When calculating forces and stress with self-consistency, we generally use the differentials of the DFT energy rather than the Harris-Foulkes energy; this enables us to separate contributions that are calculated in different ways (in particular on those that are calculated on the integration grid from those that are not).
Go to top.
Neutral atom potential
In a DFT code using local orbitals as basis functions, the total energy is most conveniently written in terms of the interaction of neutral atoms: this is simply a reformulation of the total energy which, in particular, reduces the ion-ion interaction to a sum over short-range pair-wise interactions. The charge density of interest is now the difference between the total charge density and a superposition of atomic densities, notated as \(\delta n(\mathbf{r}) = n(\mathbf{r}) - \sum_i n_i(\mathbf{r})\) for atomic densities \(n_i(\mathbf{r})\). We write:
where the second term is defined as:
The final term, the screened ion-ion interaction, is short-ranged, and written as:
where \(V_{Ha,i}(\mathbf{r})\) is the Hartree potential from the atomic density \(n_i(\mathbf{r})\). We define the neutral atom potential for an atom as \(V_{NA,i}(\mathbf{r}) = V_{L,i}(\mathbf{r}) + V_{Ha,i}(\mathbf{r})\), combining the local potential and the Hartree potential for the atomic density; the overall neutral atom potential is given as the sum over the atomic densities, \(V_{NA}(\mathbf{r}) = \sum_i V_{NA,i}(\mathbf{r})\). If we write the pseudo-atomic density as \(n_{PAD}(\mathbf{r}) = \sum_i n_i(\mathbf{r})\) then we can also write \(V_{NA}(\mathbf{r}) = V_L(\mathbf{r}) + V_{Ha, PAD}(\mathbf{r})\).
In this case, we can write the Harris-Foulkes energy as:
with the Hamiltonian defined as:
where \(V_{\delta Ha}(\mathbf{r}) = \int d\mathbf{r^\prime} \delta n(\mathbf{r^\prime})/\mid \mathbf{r} - \mathbf{r}^\prime\mid\). Accordingly, the double counting Hartree correction term has to change:
Go to top.
Non-self-consistent calculations
In non-self-consistent calculations, we use the Harris-Foulkes functional, along with a reasonable guess for the input density, which is normally taken as the superposition of atomic densities, \(n_{in}(\mathbf{r})\) and write:
Notice that we effectively have two densities being used here: \(n_{in}\) (which is normally the superposition of atomic densities used in the neutral atom case) and and effective output density, \(n_{out} = \sum_{ij} \phi_i K_{ij} \phi_j\) which comes from the band energy (first term); this complicates the calculation of forces and stress compared to the self-consistent case, as we have to consider contributions from both densities.
For the neutral atom potential, \(\delta n(\mathbf{r}) = 0\) by definition, which also means that \(E_{\delta Ha} = 0\) and \(\Delta E_{Ha}=0\).
Go to top.
Partial core corrections
Also known as non-linear core corrections, partial core corrections (PCC) [EFS1] add a model core charge to the pseudopotential to allow for the non-linear exchange-correlation interation between core and valence charge (which is linearised in standard pseudopotentials); this generally improves the accuracy of the pseudopotential. The exchange-correlation potential is evaluated in terms of the combined charge density, \(n_v(\mathbf{r}) + n_c(\mathbf{r})\) where the valence charge is input or output charge density defined above: \(V_{XC}\left[ n_v + n_c \right]\). The exchange-correlation energy becomes:
Once this change to the charge density has been made, there is no change to the DFT energy. However, the double counting term for Harris-Foulkes needs redefining, since XC contribution to the band energy is \(2Tr[KV_{XC}] = \int d\mathbf{r} n_v(\mathbf{r}) V_{XC}[n_v + n_c]\). We write:
There is an extra factor of \(\int d\mathbf{r} n_c(\mathbf{r})\epsilon_{XC}[n_v + n_c]\) over and above the usual term.
Go to top.
Forces and Stresses
It is important that the forces and stresses be the exact derivatives of the energy, for consistency. In particular, this means that as the energy is calculated in different ways for different contributions, the force or stress contribution must be calculated in the same way.
Go to top.
Forces
Forces are defined as the change in energy with respect to atomic positions; as the basis functions move with the atoms, these changes will also include Pulay terms. The forces found in Conquest are documented extensively elsewhere [EFS2, EFS3] though the changes needed to account for PCC, particularly in the non-self-consistent case, have not been published and are given here for completeness. As well as the Hellmann-Feynman forces (which come from the movement of the local and non-local pseudopotentials with the atoms) we define Pulay forces (divided into two parts, known as phi-Pulay which come from changes in the Hamiltonian matrix, and S-Pulay, which come from changes in the overlap matrix; the phi-Pulay forces are calculated in three contributions, which depend on how the respective parts of the Hamiltonian matrix are calculated: the kinetic energy; the non-local pseudopotential; and the remaining terms which are all found on the integration grid). The ion-ion interactions also contribute forces.
The inclusion of PCC adds an extra term to the forces in all calculations, which comes from the change of the core density as the atoms move; the force on atom \(i\) is given as:
If the non-self-consistent formalism is used, then a further term is added (the non-self-consistent force changes) to include the gradient of the core charge. The non-self-consistent force is now written as:
where \(V^\prime_{XC}\) is the derivative of the exchange-correlation potential with respect to charge density.
Go to top.
Stress
The stress includes all contributions to the change of energy with the lattice constants; the calculation of stress in Conquest is documented in a paper being prepared for publication, but we give a brief overview here. As Conquest uses orthorhombic cells, only the diagonal stress components (\(\sigma_{\alpha\alpha}\)) are calculated.
In most cases, forces also contribute to the stress; it is easy to show that the stress contribution is given by:
where \(R_{i\alpha}\) is the position of the atom. As well as these contributions, there are more subtle terms. Any energies calculated on the grid will contribute to the stress as the integration grid changes with cell size (the stress is simply the energy calculated), and the Hartree potential contributes a term related to the change in the reciprocal lattice vectors (as it is calculated by Fourier transforming the charge density). If the exchange-correlation functional is a GGA functional, then a further term coming from the change of the gradient of the density with the cell size arises. (For non-self-consistent calculations this leads to some complications, as this term technically requires both input and output densities; at present, we approximate this as a mixture of the term calculated with input density and the term calculated with output density; the proportion can be adjusted using the parameter General.MixXCGGAInOut
documented in the Advanced and obscure tags section of the manual, though we do not recommend changing it.)
Go to top.
Steven G. Louie, Sverre Froyen, and Marvin L. Cohen. Nonlinear ionic pseudopotentials in spin-density-functional calculations. Phys. Rev. B, 26:1738–1742, Aug 1982. doi:10.1103/PhysRevB.26.1738.
T. Miyazaki, D. R. Bowler, R. Choudhury, and M. J. Gillan. Atomic force algorithms in density functional theory electronic-structure techniques based on local orbitals. J. Chem. Phys., 121:6186, 2004. doi:10.1063/1.1787832.
Antonio S Torralba, David R Bowler, Tsuyoshi Miyazaki, and Michael J Gillan. Non-self-consistent Density-Functional Theory Exchange-Correlation Forces for GGA Functionals. J. Chem. Theory Comput., 5:1499, 2009. doi:10.1021/ct8005425.
Go to top.
Structural relaxation: Theory
Structural relaxation involves optimisation of the ionic coordinates, optimisation of the simulation cell, or both, with respect to the DFT total energy or the enthalpy if the cell is not fixed.
Ionic relaxation
L-BFGS:
To be written…
Conjugate gradients
The most naive geometry optimisation algorithm is steepest descent: we calculate the gradient of the DFT total energy (i.e. the force) and propagate the system in the direction of the steepest gradient (the direction of the force vector) until the energy stops decreasing. We choose the direction (largest gradient in this case) and perform a line search. This will be sufficient if the potential energy surface is well-behaved, but in most cases convergence will require many iterations. Conjugate gradients is a well-established method the improves upon steepest descent in the choice of search direction. Without going into too much detail, we choose a new search direction that is orthogonal to all previous search directions using the conjugacy ratio \(\beta\). At iteration \(n\), it is given by,
This is the Fletcher-Reeves formulation; note that \(\beta_0 = 0\). We can then construct the search direction at step \(n\), \(D_n\),
and peform the line minimisation in this direction. This process is repeated until the maximum force component is below some threshold.
Go to top.
Quenched MD
The system is propagated in the direction of steepest descent as determined by the DFT forces, and the velocity is scaled down as the system approaches its zero-temperature equilibrium configuration.
Go to top.
FIRE Quenched MD
The system is propagated using the modified equation of motion [Tb1],
which has the effect of introducing an acceleration in a direction that is steeper than the current direction of motion. If the power \(P(t) = \mathbf{F}(t)\cdot\mathbf{v}(t)\) is positive then the system is moving “downhill” on the potential energy surface, and the stopping criterion is when it becomes negative (moving “uphill”).
Go to top.
Cell optimisation
When optimising the cell with fixed fractional ionic coordinates, the same conjugate gradients method is used as above, but minimising the enthalpy with respect to the cell vectors.
Go to top.
Combined optimisation
The ionic and cell degrees of freedom can be relaxed simultaneously by combining all of their coordinates into a single vector and optimising them with respect to the enthalpy of the system. However, this atomic forces and total stresses having numerical values of the same order of magnitude, and changes in ionic coordinates and cell vectors being of the same order of magnitude. Using the method of Pfrommer et al. [Tb2], the latter can be enforced by using fractional coordinates for the ionic positions, and fractional lattice vectors of the form \(h = (1 + \epsilon)h_0\) where h is the matrix of lattice vectors, \(h_0\) is the matrix for some reference configuration and epsilon is the strain matrix. The “fractional” force on the i th atom is then \(\mathbf{F}_i = g\mathbf{f}_i\) where \(\mathbf{f}_i\) is the DFT-calculated force multiplied by the metric tensor \(g = h^Th\). The “fractional” stress is,
where \(\sigma\) is the DFT-calculated stress, \(p\) is the target pressure and \(\Omega\) is the volume. The resulting vector is optimised using the same conjugate gradients algorithm as before, minimising the enthalpy.
E. Bitzek, P. Koskinen, F. Gähler, M. Moseler, and P. Gumbsch. Structural Relaxation Made Simple. Phys. Rev. Lett., 97:2897, 2006. doi:10.1103/PhysRevLett.97.170201.
B. G. Pfrommer, M. Côté, S. Louie, and M. L. Cohen. Relaxation of Crystals with the Quasi-Newton Method. J. Comput. Phys., 131:233, 1997. doi:10.1006/jcph.1996.5612.
Go to top.
Molecular Dynamics: Theory
Microcanonical (NVE) ensemble
The Hamiltonian for the microcanonical ensemble is,
where \(\mathbf{p}_i\) and \(\mathbf{r}_i\) are the position and momentum of particle \(i\) and \(U\) is the DFT total (potential) energy. Hamilton’s equations can be solved to give the following equations of motion:
In order to construct a time-reversible algorithm from these equations, the Liouvillian formulation is employed [Ta1] (trivially, in this case). The Liouville operator \(L\) can be defined in terms of position and momentum components:
The Liouvillian can be used to construct the classical propagator, which relates the state \(f\) of the system at time 0 to its state at time \(t\):
Taking the individual position and momentum parts of the Liouvillian \(L_r\) and \(L_p\), it can be shown that applying it to the state \(f\) result in a simple linear shift in coordinates and a simple linear shift in momentum respectively:
However, we cannot simply replace \(e^{iLt}\) with \(e^{iL_rt}\) because \(iL_r\) and \(iL_p\) are non-commuting operators, so we must employ the Trotter-Suzuki identity:
Thus for a small enough time step \(\Delta t = t/P\) and to first order, a discrete time step corresponds to the application of the discrete time propagator \(G\),
which can be shown to be unitary and therefore time-reversible. Applying the operators \(U\) in the sequence determined by the Trotter decomposition generates the velocity Verlet algorithm, which is used to integrate microcanonical molecular dynamics in CONQUEST. For a detailed derivation of the algorithm, refer to Frenkel & Smit [Ta1].
Go to top.
Extended Lagrangian Born-Oppenheimer MD (XL-BOMD)
If the electronic density from the previous ionic step is used as an initila guess for the next SCF cycle, a problem arises because this process breaks the time-reversibility of the dynamics. This is manifested as a gradual drift in the total energy in the case of a NVE simulation, or the conserved quantity in the case of non-Hamiltonian dynamics. The solution proposed by Niklasson [Ta2, Ta3] is to introduce auxilliary electronic degrees of freedom into the Lagrangian, which can be propagated via time-reversible integrators.
The extended Lagrangian used in CONQUEST is [Ta4],
where \(\mathbf{X}\) is a sparse matrix with the same range as \(\mathbf{LS}\), \(\mu\) is the fictitious electron mass and \(\omega\) is the curvature of the auxiliary harmonic potential. The Euler-Lagrange equations of motion are then,
The first of these is simply Newton’s second law, and the velocity update equation of motion in the microcanonical ensemble. The second can be integrated using a time-reversible algorithm, the velocity Verlet scheme in the case of CONQUEST [Ta4]:
i.e. the trajectory of \(\mathbf{X}(t)\) is time-reversible, and evolves in a harmonic potential centred on the ground state density \(\mathbf{L}(t)\mathbf{S}(t)\). The matrix \(\mathbf{XS}^{-1}\) is a good guess for the \(\mathbf{L}\) matrix in the Order(N) scheme.
Despite the time-reversitibility, the \(\mathbf{X}\) matrix tends in practice to gradually drift from the harmonic centre over time, increasing the number of SCF iterations required to reach the minimum over the course of the simulation. To remove such numerical errors, the final dissipative term is included, and is found to have a minimal effect on the time-reversibility. We note that since the auxiliary variable \(X\) is used to generate an intial guess for the SCF process, it does not appear in the conserved (pseudo-Hamiltonian) quantity for the dynamics.
Go to top.
Non-Hamiltonian dynamics
Extended system method
Hamiltonian dynamics generally describes systems that are isolated from their surroundings, but in the canonical and isobaric-isothermal ensembles, we need to couple the system to an external heat bath and/or stress. It is possible to model such systems by positing a set of equations of non-Hamiltonian equations of motion, and proving that they generate the correct statistical ensemble [Ta5]. This is the extended system approach: we modify the Hamiltonian to include the thermostat and/or barostat degrees of freedom, derive the (pseudo-) Hamiltonian equations of motion, and demostrate that the correct phase space distribution for the ensemble is recovered.
Go to top.
Canonical (NVT) ensemble
In the Nose-Hoover formulation [Ta6, Ta7], the Hamiltonian for a system in the canonical ensemble can be written,
where \(\mathbf{r}_i\) and \(\mathbf{\dot{r}_i}\) are respectively the position and velocity of particle \(i\), \(U\) is the potential energy, in this case the DFT total energy, \(s\) is a dimensionless quantity that can be interpreted post-hoc as a time step scaling factor, \(Q\) is the fictitious mass of the heat bath and \(n_f\) is the number of ionic degrees of freedom. Hamilton’s equations can be solved to generate the Nose-Hoover equations of motion. However Martyna et al. demonstrate that this method does not generate an ergodic trajectory, and proposed an alternative formulation [Ta8] in which the temperature is controlled by a chain of \(M\) coupled thermostats of mass \(Q_k\), notional position \(\eta_k\) and conjugate momentum \(p_{\eta_k}\):
The Liouvillian for these equations of motion can be non-uniquely decomposed into components of ionic position (\(iL_r\)) and momentum (\(iL_p\)) as in the microcanonical case, the extended Lagrangian (\(iL_\mathrm{XL}\), and a Nose-Hoover chain component (\(iL_\mathrm{NHC}\))
which is directly translated into an algorithm with the Trotter-Suzuki expansion,
This is recognisable as the velocity Verlet algorithm with extended Lagrangian integration which can be reduced to a single step, as described in Extended Lagrangian Born-Oppenheimer MD (XL-BOMD), with a half time step integration of the Nose-Hoover chain equations of motion before and after. For full details of the integration scheme, see Hirakawa et al. [Ta9].
Go to top.
Isobaric-Isothermal (NPT) ensemble
The Parinello-Rahman equations of motion [Ta10] extend the fixed cell equations of motion to include the cell degrees of freedom in the extended system approach. We use the Martyna-Tobias-Tuckerman-Klein modification [Ta11], which couples the variable cell equations of motion to a Nose-Hoover chain the themrostat the system, recovering the isobaric-isothermal (NPT) ensemble. For an unconstrained cell (i.e. the lattice vectors can change freely), the equations of motion are,
Here, \(\mathbf{r}_i\), \(\mathbf{p}_i\) and \(m_i\) are the position, momentum and mass of particle \(i\) respectively, \(\xi\), \(p_\xi\) and \(Q\) are the position, momentum and mass of the thermostat, and \(\mathbf{h}\), \(\mathbf{p}_g\) and \(W_g\) are the matrix of lattice vectors, matrix of cell velocities and cell mass respectively. Note that these equations only include one Nose-Hoover thermostat for simplicity. Conquest uses the Shinoda-Shiga-Mikami splitting of the Liouvillian [Ta12] to propagate the system. The Liouvillian is decomposed as,
which can be further split,
Using Liouville’s theorem, we have,
Here we use \(M\) heat baths in a Nose-Hoover chain. The Trotter-Suzuki expansion is,
The Liouvillian for the heat baths can be further expanded:
Finally, expanding the first propagator in the previous expression, we have,
These expressions are directly translated into the integration algorithm.
Go to top.
Weak coupling thermostat/barostat
Instead of modifying the Hamiltonian, the Berendsen-type weak coupling method [Ta13] involves coupling the ionic degrees of freedom to a an external temperature and/or pressure bath via “the principle of least local perturbation consistent with the required global coupling.” Thermostatting is acheived via a Langevin-type equation of motion, in which the system is globally coupled to a heat bath and subjected to random noise:
where \(\gamma\) is a global friction constant chosen to be the same for all particles. This can be acheived in practice by rescaling the velocities \(\mathbf{v}_i \rightarrow \lambda\mathbf{v}_i\), where \(\lambda\) is,
A similar argument can be applied for weak coupling to an external pressure bath. In the isobaric-isoenthalpic ensemble, the velocity of the particles can be expressed,
i.e. the fractional coordinates are scaled by a factor determined by the difference between the internal and external pressures, the isothermal compressibility \(\beta\) and a pressure coupling time constant $tau_P$. In the isotropic case, the cell scaling factor \(\mu\) can be expressed,
where the compressibility is absorbed into the time time constant \(\tau_P\). Allowing for fluctuations of all cell degrees of freedom, the scaling factor becomes,
While trivial to implement and in general stable, the weak-coupling method does not recover the correct phase space distribution for the canonical or isobaric-isothermal ensembles, for which the extended system method is required. It is no longer supported in CONQUEST, but the concepts are useful.
Go to top.
Stochastic velocity rescaling
Stochastic velocity rescaling (SVR) [Ta14] is a modification of the weak coupling method, in which a correctly constructed random force is added to enforce the correct NVT (or NPT) phase space distribution. The kinetic energy is rescaled such that the change in kinetic energy between thermostatting steps is,
where \(\bar{K}\) is the target kinetic energy (external temperature), \(dt\) is the time step, \(\tau\) is the time scale of the thermostat, \(N_f\) is the number of degrees of freedom and \(dW\) is a Wiener process. Practically, the particle velocities are rescaled by a factor of \(\alpha\), defined via,
Where \(R_i\) is a set of \(N_f\) normally distributed random numbers with unitary variance. This method can be applied to thermostat the NPT ensemble by barostatting the system with the Parinello-Rahman method, and using the above expressions, but with additional \(R_i\)’s for the cell degrees of freedom, and thermostatting the cell velocities as well as the particle velocities [Ta15].
D. Frenkel and B. Smit. Understanding molecular simulation: from algorithms to application. Academic Press, 2002.
A. M. N. Niklasson. Extended Born-Oppenheimer Molecular Dynamics. Phys. Rev. Lett., 100:123004, 2008. doi:10.1103/PhysRevLett.100.123004.
A. M. N. Niklasson and M. J. Cawkwell. Generalized extended Lagrangian Born-Oppenheimer molecular dynamics. J. Chem. Phys., 141:164123, 2014. doi:10.1063/1.4898803.
M. Arita, D. R. Bowler, and T. Miyazaki. Stable and Efficient Linear Scaling First-Principles Molecular Dynamics for 10000+ Atoms. J. Chem. Theor. Comput., 10:5419, 2014. doi:10.1021/ct500847y.
M. E. Tuckerman. Statistical mechanics: theory and molecular simulations. Oxford Graduate Texts, 2010.
S. Nosé. A unified formulation of the constant temperature molecular dynamics methods. J. Chem. Phys., 81:511, 1984. doi:10.1063/1.447334.
W. G. Hoover. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A, 31:1695, 1985. doi:10.1103/PhysRevA.31.1695.
G. J. Martyna, M. L. Klein, and M. Tuckerman. Nosé–hoover chains: the canonical ensemble via continuous dynamics. J. Chem. Phys., 97:2635, 1992. doi:10.1063/1.463940.
T. Hirakawa, T. Suzuki, D. R. Bowler, and T. Miyazaki. Canonical-ensemble extended lagrangian born-oppenheimer molecular dynamics for the linear scaling density functional theory. J. Phys.: Condens. Matter, 29:405901, 2017. doi:10.1088/1361-648X/aa810d.
M. Parrinello and A. Rahman. Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys., 52:7182–7190, December 1981. doi:10.1063/1.328693.
G. J. Martyna, M. E. Tuckerman, D. J. Tobias, and M. L. Klein. Explicit reversible integrators for extended systems dynamics. Mol. Phys., 87:1117, 1996. doi:10.1080/002689799163235.
W. Shinoda, M. Shiga, and M. Mikami. Rapid estimation of elastic constants by molecular dynamics simulation under constant stress. Phys. Rev. B, 69:4396, 2004. doi:10.1103/PhysRevB.69.134103.
H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. Dinola, and J. R. Haak. Molecular dynamics with coupling to an external bath. J. Chem. Phys., 81:3684, 1984. doi:10.1063/1.448118.
G. Bussi, D. Donadio, and M. Parrinello. Canonical sampling through velocity rescaling. J. Chem. Phys., 126:014101, 2007. doi:10.1063/1.2408420.
G. Bussi, T. Zykova-Timan, and M. Parrinello. Isothermal-isobaric molecular dynamics using stochastic velocity rescaling. J. Chem. Phys., 130:074101, 2009. doi:10.1063/1.3073889.
Go to top.
Get in touch
If you have suggestions for developing the code, please use GitHub issues. The developers cannot guarantee to offer support, though we will try to help.
Report bugs, or suggest features on GitHub issues. View the source code on the main GitHub page.
You can ask for help and discuss any problems that you may have on the Conquest mailing list (to register for this list, please send an email to mlsystem@ml.nims.go.jp with the subject “sub conquest-user”, though please note that, for a little while, some of the system emails may be in Japanese; you will receive a confirmation email to which you should simply reply without adding any text).
Licence
CONQUEST is available freely under the open source MIT Licence. We ask that you acknowledge use of the code by citing appropriate papers, which will be given in the output file (a BiBTeX file containing these references is also output). The key CONQUEST references are:
A. Nakata, J. S. Baker, S. Y. Mujahed, J. T. L. Poulton, S. Arapan, J. Lin, Z. Raza, S. Yadav, L. Truflandier, T. Miyazaki, and D. R. Bowler, J. Chem. Phys. 152, 164112 (2020) DOI:10.1063/5.0005074
T. Miyazaki, D. R. Bowler, R. Choudhury and M. J. Gillan, J. Chem. Phys. 121, 6186–6194 (2004) DOI:10.1063/1.1787832
D. R. Bowler, T. Miyazaki and M. J. Gillan, J. Phys. Condens. Matter 14, 2781–2798 (2002) DOI:10.1088/0953-8984/14/11/303