Structural optimization using forces and stresses

This tutorial provides an overview of the methods available in Siesta for structural optimization, which are quite varied in their nature, scope, and sophistication.

Note

For background, you can refer to these slide presentations from past Siesta Schools:

Most DFT codes use the Hellman-Feynman theorem to compute forces and stresses directly, without having to perform several energy calculations and use finite differences. This affords tremendous gains in efficiency. With forces and stresses at hand given the structure, one can implement algorithms that will search for the optimal structure with zero forces and stresses.

Note

This is a good time to review the issues related to the use of a real-space grid, and how to get appropriately accurate forces and stresses.

We have already seen geometry relaxation in action in the context of the CH4 molecule in A first encounter with Siesta. We will re-take molecules further below, when we discuss constraints, but now we turn to examples with crystals.

Hint

Enter directory SiH

We will use the (slightly artificial) example of a cubic box of 64-atoms of Si with an extra H atom placed in the middle of a Si-Si bond. That is not quite optimal, and the H atom ‘pushes out’ its neighbors to lower the energy and minimize the forces.

For the purposes of running this example quickly, we are using a SZ basis set and a very low cutoff. You might want to move beyond this later, but there should not be qualitative changes in the results.

We look first at the options for relaxation:

MD.TypeOfRun         CG
MD.NumCGsteps        50
MD.MaxForceTol       0.05 eV/Ang

CG stands for “conjugate gradients”, and is one of the standard optimizers within Siesta.

For this example the program will run for around 15 geometry iterations (that is, steps in which the structure is modified in response to the forces, until the cycle converges). The tolerance for convergence is given by the last line above (note the explicit units), and is slighty above the default in Siesta (0.04 eV/Ang).

You might want to see the evolution of the total energy of the system during the run. You can scroll through the file or, as a shortcut, use the command (assuming you have redirected the output to file OUT:

grep enth OUT

There is a gain of more than 5 eV in the relaxation.

To check the enviroment of the H atom before and after the relaxation, you can look in the .BONDS and .BONDS_FINAL files:

tail *.BONDS        # distances to the neighbors of H
                    # at the start

tail *.BONDS_FINAL  # distances to the neighbors of H
                    # at the end

You can also use a visualizer program that can read the .xyz file produced by the option write-coor-xmol T in the fdf file.

It is possible to use a variety of optimization engines within Siesta. Above we used MD.TypeOfRun CG to select the conjugate-gradients engine, but one could use also MD.TypeOfRun Broyden, or MD.TypeOfRun FIRE to employ other algorithms, which might be more efficient than CG. (Test this by replacing the option and starting the relaxation again – you should see, for example, that the Broyden optimizer needs fewer steps (about half!) to reach convergence.)

These optimization engines are coded in Siesta itself. It is possible to use the Lua scripting engine embedded in Siesta to access other algorithms. In particular, a version of the very efficient LBFGS algorithm can be used for structural optimization.

Variable-cell optimization

In the example above the atoms are moved in response to the forces, but the lattice vectors stay fixed. If you have scrolled through the output file you might have seen lines mentioning “stress”. For example:

Stress tensor Voigt[x,y,z,yz,xz,xy] (kbar):      -70.56      -70.56      -70.56        0.37        0.37        0.37

A non-zero stress means that the lattice vectors are not optimal, and the energy can be lowered by optimizing them. The relaxation algorithm uses now two sets of variables: atomic coordinates and lattice vectors, and moves them in response to the forces and the stresess, until these are below a set tolerance.

Variable-cell optimization is non-trivial to implement or run properly, as the forces and the stresses are not only dependent on the coordinates and lattice vectors, respectively. Both sets are coupled. Besides, the dimensions (energy/length and energy/volume, respectively for forces and stresses) and the ‘spring constants’ that determine the changes in them in response to distortions are of course different.

Hint

Enter directory varcell_cg

The example we will use is an 8-atom Si cell (the ‘conventional cubic cell’). Note the options:

MD.TypeOfRun          CG
MD.NumCGSteps         100
MD.VariableCell       T
MD.MaxForceTol        0.1 eV/Ang
MD.MaxStressTol       0.1 GPa
MD.TargetPressure     0.0 GPa

As expected, the tolerances have different physical dimensions. The last line tells the program to aim for a zero (diagonal) stress (in practice, “atmospheric pressure”), but it can be changed to obtain the optimum geometry of the system under applied pressure (hydrostatic in this case, but an arbitrary tensor target can be also specified for non-hydrostatic conditions – see the manual).

The example is set-up to start with a structure determined by:

LatticeConstant     5.535 Ang
%block LatticeVectors
 1.150  0.200  0.000
 0.000  1.050  0.000
-0.100  0.000  0.900
%endblock LatticeVectors

We see that the first two cell vectors are too large, so the system is under tensile stress along the x and y directions. Conversely, the third cell vector is too small (compressive stress). In addition, there are off-diagonal components of the strain, leading to ‘shear’ components of the stress. Note the signs and sizes of these initial stresses, and watch how they move towards zero. (You can do this by grep oigt OUT, where OUT is the output file you have chosen.) Also, check the evolution of the energy (grep enth OUT)

There are a number of things to note:

  • The stresses do not decrease monotonically to zero.

  • The energy does not decrease monotonically either.

  • The final lattice vectors look “funny” when one looks at their cartesian components (for example, in file si8.STRUCT_OUT), but their modules are roughly the same and the angles between them are very close to 90 degrees (grep for modules and angles in the output file). The cell might have rotated during the process of relaxation, but it is basically cubic at the end.

  • The atomic coordinates (in fractional form) are very close to their initial values, which are the standard sites in the Si diamond structure (in the conventional cell). This is to be expected, since these are high-symmetry positions.

Some of these might be related to, or made worse by, the ridiculously low mesh-cutoff chosen (30 Ry) and by the small basis set (try improving these, but note the increased cpu time).

Relaxation with “quenched” molecular dynamics

There is an alternative relaxation method that uses a physically motivated scheme, rather than a purely mathematical search for a ‘zero forces and stresses’ configuration. Imagine that we perform a molecular dynamics simulation in which, rather than relaxing, we move the atoms, and the cell vectors, according to the (classical) equations of motion, using the forces and stresses. For more information about this, see this tutorial. The trick in this case is that, every time an atom senses a force ‘opposite’ its velocity (in the sense that their scalar product is negative), the velocity is set to zero. This roughly corresponds to the idea: “since the atom seems to be moving away from its equilibrium point, we rather stop it”. The same can be done with strains and stresses in the case of variable cell.

Hint

Enter the directory varcell_md

Look now at the “relaxation section” of the si8.fdf file:

MD.TypeOfRun          ParrinelloRahman
MD.InitialTimeStep    1
MD.FinalTimeStep      200
MD.LengthTimeStep           3.0 fs
MD.ParrinelloRahmanMass  10.0 Ry*fs**2
MD.Quench             T

The ParrinelloRahman scheme is a combined atoms+cell microcanonical scheme (see MD tutorial). We allow it to run for up to 200 steps, with a time-step of three femtoseconds. Note also the appearance of a “mass” with the dimensions of energy*time^2: this is used to homogeneize the dynamics of the system, which has to deal, as we indicated earlier, with fundamentally different sets of variables. The final line request the “quenching”.

If you now run the example, you will notice that it converges quite nicely, with monotonic decrease in the energy and cleaner evolution of the stresses (even if not monotonic). This method is always quite robust, and in this particular case of variable-cell, particularly efficient. It can generally be counted on to bring systems closer to the optimal structure, and the final relaxation can be done with a faster method.

Note

We can even start with some extra kinetic energy, in the form of some ‘starting temperature’, to wiggle things around and free the system from any undesired local minimum.

Constraints

Our next system is a very simple model of the H-terminated (100) as-cut surface of Si. It is a periodically-repeated slab, with three layers of Si atoms (six atoms in total), and four H atoms. We want to know how the top-most section relaxes, while maintaining the “bulk-like” bottom layers fixed to simulate the connection to the bulk below.

Hint

Enter the si100_constrained directory

Note the block:

%block GeometryConstraints
      position from 1 to 4
%endblock GeometryConstraints

which requests that the positions of the first four atoms (those at the bottom of the slab) are kept fixed.

The system converges (with the Broyden algorithm) in about 15 steps. During the process, forces on all atoms are computed, but only those on non-constrained atoms enter into the check for convergence (the maximum absolute value of any component of these forces is what grep constrained OUT prints).

There are many options for the specification of relaxation constraints in Siesta, including features such as fixing a group of atoms (‘molecule’) to move rigidly together, constraining cell vector sizes or angles, etc. You should check the manual to get more information.

In the next section we will explore another way to express constraints, by using reduced coordinates.

Z-matrix constraints

Hint

Enter the h2o_zmatrix directory

Sometimes the important structural degrees of freedom that we want to optimize are not easily represented in cartesian coordinates. In molecules, for example, bond lengths, angles, and torsion angles are much more relevant than particular values of the cartesian coordinates. For many years, chemists have used more appropriate ways to represent molecular structure, and the Zmatrix is one of them. Consider this block in file h2oZ.fdf:

%block Zmatrix
molecule_cartesian
  1 0 0 0   0.0 0.0 0.0 0 0 0
  2 1 0 0   HO1 90.0 37.743919 1 0 0
  2 1 2 0   HO2 HOH 90.0 1 1 0
variables
    HO1 1.0
    HO2 1.0
    HOH 106.0
%endblock Zmatrix

which represents the structure of a water molecule by giving the cartesian coordinates of the oxygen atom (placed at the origin), but using essentially bond-lengths (HO1 and HO2), and the HOH angle for the H atoms, instead of coordinates. (There are other pieces of data that can be explained by looking in the manual.).

Furthermore, these symbols represent variables, which can vary during a relaxation. We have set initially, for example, the bond-length to the bball-park value of 1 Ang, and the HOH angle to 106 degrees. If you run the example you will see how this variables are changed until relaxation within the tolerances is achieved. The tolerances themselves have a new, more appropriate form:

ZM.ForceTolLength 0.04 eV/Ang
ZM.ForceTolAngle 0.0001 eV/deg

You might want to play further with this example in several directions:

  • Fix the HOH bond angle. For this, introduce a new section constants: in the block (see the manual) and place it there.

  • Compare the results to experimental data. Maybe you need to use a GGA functional for better results. Try to get (PSML) pseudopotentials from Pseudo-Dojo for this.

  • You can use an extendend Zmatrix format to study molecules near surfaces. See the manual for an example.