Monthly Archives: February 2018

Comparing Molecular Geometry Optimization with Localized and non-Localized Basis Sets

Purpose of Calculation

Geometric optimization of a unit cell is useful for finding the minimum energy configuration of a molecule or solid without needing to manually change atom positions. Through numeric minimization algorithms, the proper lattice parameters and atom positions can be obtained. For molecules, this can also find the minimum energy bond angles and bond lengths. However, the different contexts between solids and molecules can present very different challenges. These calculations will compare the convergence and speed of the geometry optimization calculations of water and hydrogen cyanide (HCN), using two different basis sets: Castep’s [1] plane wave basis set and DMol3’s calculated-on-the-fly localized basis set [2].

Figure 1: Optimized geometry of a water molecule

Figure 2: Optimized geometry of a hydrogen cyanide molecule in the H-C-N isomer

Figure 3: Optimized geometry of a hydrogen cyanide molecule in the H-N-C isomer

Calculation Methodology

The bond angle and bond lengths of water and hydrogen cyanide are calculated using two different basis sets.

All calculations are performed with the Perdew-Burke-Ernzerhof (PBE) [3] exchange-correlation functional, a Generalized Gradient Approximation (GGA) functional. The Koelling-Harmon relativistic treatment was used for atomic solutions [4] in the Castep calculations for the plane wave basis set. The on-the-fly basis set is calculated by generating a radial function calculated by solving the density equations on the fly at three hundred points on a logarithmic radial mesh for each atom. This radial function is then multiplied by appropriate spherical harmonics for each atom.

Convergence Calculations

Calculations to check the convergence of the minimum energy output with respect to the energy cutoff, the k-point mesh and the unit cell size were performed for the Castep calculations. It is necessary to check the convergence with respect to the unit cell size because the plane wave basis fills the entire cell. Due to the computational assumption of repeating cells, each unit cell must be large enough so that each molecule does not interact with those in adjacent cells. Convergence tests must be used to idea what is a sufficiently large cell. These tests are shown in figures 4-9 for both water and hydrogen cyanide.

 

Fig. 4 Convergence of Castep with respect to Cutoff Energy for water

Fig. 5 Convergence of Castep with respect to k-point number for water

Fig. 6 Convergence of Castep with respect to unit cell size for water

Fig. 7 Convergence of Castep with respect to Cutoff Energy for hydrogen cyanide

Fig. 8 Convergence of Castep with respect to k-point number for hydrogen cyanide

Fig. 9 Convergence of Castep with respect to unit cell size for hydrogen cyanide

For DMol3, we do not need to converge the unit cell size. The generated basis sets are highly localized, and we instead test the convergence of the maximum radius of the radial functions, referred to as the cutoff radius. The cutoff radius must be large enough that each atom “sees” the other atoms during all of the geometry optimization, but can still be smaller enough that the basis does not extend as far as the unit cell extends in the Castep calculations. As a result, we check the convergence of energy calculations with respect to cutoff radius and k-point mesh. These convergence tests are shown in figures 10-13 for both water and hydrogen cyanide.

Fig. 10 Convergence of DMol3 with respect to k-point number for water

Fig. 11 Convergence of DMol3 with respect to cutoff radius for water

Fig. 12 Convergence of DMol3 with respect to k-point number for hydrogen cyanide

Fig. 13 Convergence of DMol3 with respect to cutoff radius for hydrogen cyanide

All parameters were converged to within 0.01 eV

Calculation Results

The key metric for comparing the two sets of results is the computational time for both the geometry optimization and the convergence tests. Given an arbitrarily long time, either basis set can give an equally accurate and precise calculation of the bond angles and bond lengths. The important comparison is how quickly the computations can be performed using each basis set to within reasonable agreement with each other.

Both basis sets agree that there is one minimum for the water molecule: a planar bent structure, with bond angle of approximately 104 degrees and a bond length of approximately 0.97 angstroms. Both basis sets also report two isomers of hydrogen cyanide, the H-C-N configuration, and the C-N-H configuration, both with a linear orientation and bond angle of approximately 180 degrees. Disagreement on the exact bond angles and lengths is minor, but does occur. Exact results are included in table one. By better converging the minimum energy, a geometry optimization with a stricter cutoff could be run in order to lessen such disagreement. This, however, is not the purpose of this example case.

BondCastep Bond Length (Angstroms)Castep Bond Angle (Degrees)DMol3 Bond Length (Angstroms)DMol3 Bond Angle (Degrees)
H-O0.9700.971
H-O-H103.708103.708
H-C (H-C-N Isomer)1.0751.076
C-N (H-C-N Isomer)1.1591.164
H-C-N (H-C-N Isomer)175.507179.406
H-N (H-N-C Isomer)1.0061.005
N-C (H-N-C Isomer)1.1761.181
H-N-C (H-N-C Isomer)179.970179.959

The total computation time for both convergence studies and the geometry optimizations for Castep is 8486 seconds (141.43 minutes). For DMol3, the total time is 102 seconds, providing a similar degree of accuracy. This result is expected, due to the more localized basis set being favorable for describing the more local behavior of a molecule confined to a small portion of a cell. The size of the disparity between the two methods is very noteworthy, however, and shows how helpful it can be to choose the correct basis set for a give task.

References

[1]  S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. J. Probert, K. Refson, M. C. Payne, “First principles methods using CASTEP”, Zeitschrift fuer Kristallographie 220(5-6) pp. 567-570 (2005)

[2]  Delley, J., Fast Calculation of Electrostatics in Crystals and Large Molecules; Phys. Chem. 100, 6107 (1996)

[3]  John P. Perdew, Kieron Burke, and Matthias Ernzerhof, “Generalized Gradient Approximation Made Simple”, Phys. Rev. Lett. 77, 3865 – Published 28 October 1996; Erratum Phys. Rev. Lett. 78, 1396 (1997)

[4]  D D Koelling and B N Harmon 1977 J. Phys. C: Solid State Phys. 10 3107

Surface relaxiation of (0001) face in hematite

Introduction

This project aims to study the (0001) surface relaxation of hematite, including determining a suitable surface size and thickness for surface energy calculation. Density Functional Theory (DFT) calculations for hematite surface geometry optimization were performed using the Perdew-Burke-Ernzerh [1] of generalized gradient approximation for the exchange-correlation functional in the DMol3 code [2]. The Monkhorst -Pack k-point grid in this calculation is determined by the convergence of the free energy of bulk hematite. The slab thickness is chosen when the atomic layer in the center of the slab remains the same after DFT calculation.

Hematite Crystal structure

This project uses energy converged hematite (α-Fe2O3) structure with lattice parameter a=b=5.05Å, c=13.81Å, α=β=γ=90 degrees, and space group R-3c [3]. The equivalent fractional coordination in bulk hematite for iron is (0, 0, 0.353694), for oxygen is (0.3059, 0, 0.25). Iron occupies two-thirds of slightly distorted octahedron sites, and oxygen locates at the tetrahedron site. The Fe-O3-Fe unit repeated along the z-axis and stacked oxygens form a hexagonal close-packed (HCP) structure (Figure 1).

Figure 1. The structure of hematite, with its space group R-3c. Oxygens (red) stacked hexagonal close-packed, and irons (blue) locate at two-thirds of the octahedron site.

Construction of (0001) Surface

Former research shows that surface (0001) in hematite react to water and metal legends actively. The (0001) surface of the hematite model was built with half-infinite slabs, ten atomic layers, and 10Å vacuum. Choosing symmetric slab reduces the generated dipole between uppermost and button layer. The thickness of the slab should be large enough to ensure the surface is enough for relaxation. In this calculation, we use ten atomic layers to relax the surface atom. The distances between O-Fe and Fe-Fe atomic layers are shown in Figure 2. Due to the distorted octahedra site in hematite structure, the irons are not in the same atomic plane and thus regarded as in two different planes with their distance of 0.607Å.

Figure 2. Ten atomic layers present with (0001) cleaved surface, blue is iron and red is oxygen.

 

(0001) Surface relaxation result

The slab thickness can only be chosen if the atomic positions at the center of the slabs retain the same with bulk coordinates. A DFT-optimized 1*1 surface geometry optimization was calculated initially using a 5*5*1 Monkhorst-Pack k-point grind and 4.5Å global orbital cutoff [4]. The core treatment for iron and oxygen atoms is applied to all electrons. The energy convergence and SCF tolerance are 1.0e-5 Hartree eV and 1.0e-6 eV separately. The maximum force applied on each atom is 0.002 Ha/Å, and maximum displacement for each atom is 0.05 Å. Figure 3 shows iron and oxygens coordination after relaxation. The relaxed layer-1 oxygen anions have moved 2.21% to the positive z-direction from their bulk positions; the relaxed layer-2 iron has moved 1.31% to the negative z-direction from the bulk positions. The distance of first two layers and last two layers showed a contraction from 0.847 Å  to 0.688 Å (Figure3). Moreover, the oxygen in the first layer moved not only along z axis but also along x and y axis to relax the surface energy (Figure4).

Figure 3. DMol3 calculation with K-point set 5*5*1

Figure 4. A comparison of first layer oxygen original coordinates (left) and after surface relaxation (right). Oxygen in the first and last layer changed all their coordinates to relax surface energy.

The energy decreased and converged as the atoms find their preferred location (Figure 5). The  total energy of using 5*5*1 and 6*6*1 k-points is -8483.9002004eV and -8483.900402eV. This geometry optimization calculation shows the coordination of the center of the atom layer (layer 5-6, Fe-Fe layer) changed swiftly than in bulk after calculation, suggesting this model required larger suitable slab size for geometry optimization. But we get similar total energy and percent realization using 5*5*1 and 6*6*1 k-points, showing that the k-point is suitable for the geometry optimization (Table 1). 

Figure 5. Energy convergence in geometry optimization, giving a result of -8483.900402 eV using 6*6*1 k-point

Table 1. DMol3 geometry optimization result and percent relaxation of surface (0001)  with k-point 5*5*1

Conclusion

In this project, the slab was built symmetrically with a 10Å vacuum. The Brillouin zone integrated with a 5*5*1 Monkhorst-Pack k-point grid. The global orbital cutoff is set as 4.5Å. However, the slab in this calculation is not enough to relax the (0001) surface within ten atomic layers because the center layer of the atoms moved too much after DFT calculation. A good indication of surface relaxation is the atom position of the center remaining the same after relaxation while a first and last couple of layers are relaxed.

 

Reference

[1] Perdew, J. P., Burke, K., & Ernzerhof, M. (1996). Generalized gradient approximation made simple. Physical review letters, 77(18), 3865.

[2] Delley, B. (2000). From molecules to solids with the DMol 3 approach. The Journal of chemical physics, 113(18), 7756-7764.

[3] Trainor, T. P., Chaka, A. M., Eng, P. J., Newville, M., Waychunas, G. A., Catalano, J. G., & Brown Jr, G. E. (2004). Structure and reactivity of the hydrated hematite (0 0 0 1) surface. Surface Science, 573(2), 204-224.

[4] Lo, C. S., Tanwar, K. S., Chaka, A. M., & Trainor, T. P. (2007). Density functional theory study of the clean and hydrated hematite (1 1¯ 02) surfaces. Physical review B75(7), 075425.

 

 

Finding the Equilibrium Geometry of Hydrogen Cyanide and Water

In this work we use density functional theory to find the preferred geometry of molecular pairs of hydrogen cyanide (HCN) and water (H₂0). We performed these calculations for all electrons within the framework of DMol³⁽¹·²⁾ using a basis set of atomic orbitals and the PBE⁽³⁾ functional for the exchange-correlation energy. The specific basis set used was DNP 3.5 (double numerical with polarization).

Methods

As a precursor to our main result, we performed geometry optimization for isolated H₂O and then isolated HCN and compared our results to values in the literature. This served a dual purpose, not only providing us with a suitable initial geometry for each of the molecules in our main calculation, but also giving us an idea the systematic errors implicit in our method. Our calculations corroborate and are corroborated by the values reported in the literature, as may be seen in Table 1.

Table 1: Geometry of HCN and H2O

CN bond length (Å)CH bond length (Å)OH bond (Å)HOH angle
Literature1.156⁽⁴⁾1.068⁽⁴⁾0.958⁽⁵⁾104.5°⁽⁵⁾
Calculated1.161.0740.969103.64°

Table 1: Geometry of isolated HCN and H₂O

Because the geometry optimization algorithms proceed by using calculated forces to update initial atomic positions, it is critical that we choose judiciously the set of initial geometries we will optimize. Individual initial geometries must be chosen so that appreciable forces exist, and the set of initial geometries must be diverse enough to capture all local minima. Here we discuss our choice of initial geometries.

As free parameters our we may choose the three-variable locations (e.g. three Cartesian coordinates) and two-variable orientations (e.g. polar and azimuthal angles) of each molecule. This is immediately reduced by the fact that only the relative displacement and orientations of the molecules matter, and we exploit this freedom by choosing to place the carbon of HCN at the origin initially and align it with the z-axis. Furthermore, because HCN is linear, it is identical upon rotations about this axis and hence we can also take H₂O to be in the xz plane. Symmetry in reflection across the yz plane means we can further narrow the space to be considered to the portion of the xz plane with x>0. We are left with four parameters for H20, two for each of its location and its orientation.

With the bond lengths and angles fixed, we chose five initial positions for the water molecule, each with two orientations. At a distance of 2-3Å from the carbon of HCN, H₂O was placed at approximately 30°, 60°, 90°, 120° and 150° from the z axis. For each location one optimization began with the H₂O dipole either parallel or antiparallel to the x axis, with the plane defined by the three atoms of H₂O at a 45° angle to the xz plane. Two examples are shown below in Figure 1.

Figure 1 – Two example initial geometries; with C of HCN as the origin, the upper H₂O is at 30° off the z-axis (measured as the angle O-C-N) with a dipole parallel to the x-axis and the lower H₂O is at 120° with a dipole antiparallel to the x-axis

Our minimization criterion was to have the energy and displacement between successive steps vary by less than, respectively, 10⁻⁵Ha and 0.002Ha/Å. Those interested in reproducing our results should note that, with our maximum step size of 0.3Å, in a few cases (namely, those in which H₂O was initially repulsed by dipole forces) this took slightly over 50 iterations to converge.

Results

All equilibria we found fell into one of three categories. The lowest in energy had a hydrogen bond between the H of HCN and O of H₂O with an energy of -169.7358Ha; the dipole of H₂O formed an angle of approximately 30° with the axis of HCN. This is shown in Figure 2. The angle between the dipoles can be qualitatively explained by thinking of the tetrahedral geometry of oxygen’s two bonds and two lone pairs under sp³ hybridization. If H₂O were a regular tetrahedron and one lone pair were aligned with the HCN axis, some geometry leads to 35° as the expected angle, so the tetrahedral shape of a molecule with sp³-hybridized orbitals explains our results well.

In the next lowest, a hydrogen bond formed between an H of H₂O and the N of HCN, with the relevant O-H bond aligned with the linear HCN. The energy of this arrangement was -169.7337Ha. This difference is about 0.05eV, or about 660K on the Boltzmann energy scale. (There is of course an “activation energy” necessary to switch configurations which exceeds this gap, but this is not the subject of this project.)

Figure 2 – The two miminal-energy geometries of H₂O and HCN

All but one of the convergent optimizations described above ended with the lowest-energy state; only when H₂O was placed at 30° off the z-axis with its two Hs (its positive poles) facing the N of HCN (its negative pole). When the opposite orientation was assumed (i.e. O closest to N) the repulsive force so outweighed any force in the z-direction that H₂O simply drifted away from HCN. An additional calculation the dipole of H₂O orthogonal to that of HCN, with H₂O centered 25° off the z-axis, also converged to this secondary local minimum. Thus the H₂O-NCH geometry (here the atoms connected by hyphens share a hydrogen bond) is the stable energy miminum for a large majority of initial geometries, with a moderately strong but not insurmountable difference between the two minima.

Conclusion

We have determined that the most energetically favorable arrangement of a single molecule each of H₂O and HCN is that in which O in H₂O forms a hydrogen bond with H in HCN. We also showed that a second local energy minimum exists in which one of the two H in H₂O forms a hydrogen bond with N in HCN. The existence of multiple local minima suggests that a many-molecule system could have very interesting geometries – for example, one might imagine that it could more energetically favorable for some molecules to take the slightly higher-energy arrangement if it would allow more energy-lowering hydrogen bonds to occur elsewhere. The possibities could be further expanded by considering HCN’s higher-energy tautomer hydrogen isocyanide, structurally written NCH. The single-H₂O, single-HCN nature of our work elucidates the “building blocks” of such many-particle interactions and builds a good foundation for future work on increasingly complex systems.

References

[1] B. Delley, “An all‐electron numerical method for solving the local density functional for polyatomic molecules”.J. Chem. Phys. 92, 508 (1990).

[2] B. Delley, “From molecules to solids with the DMol<sup>3</sup> approach”. J. Chem. Phys. 113, 7756 (2000).

[3] J. P. Perdew, K. Burke and M. Ernzerhof, “Generalized gradient approximation made simple”. Phys. Rev. Lett. 77, 3865 (1996).

[4] L.E. Sutton, ed., “Tables of Interatomic Distances and Configuration in Molecules and Ions”. London: The Chemical Society (1958). (via www.wiredchemist.com/chemistry/data/carbon-compounds)

[5] A. G. Császár, G. Czakó, T. Furtenbacher, J. Tennyson, V. Szalay, S. V. Shirin, N. F. Zobov and O. L. Polyansky, “On equilibrium structures of the water molecule”. J. Chem. Phys., 122, 214305 (2005). (via http://www1.lsbu.ac.uk/water/ref9.html#r836)

Determination of the Optimal Structure of Water and Hydrogen Cyanide

1. Problem Description

In this project, we aim to predict the optimal structures of H2O and HCN using DFT. To calculate the energy of an isolated molecule with periodic supercells, the molecule is placed in a large enough cubic supercell to minimize the influence from period images. To perform the geometric optimization, we need to take all the internal degrees of freedom like bond lengths and bond angles into consideration with appropriate stopping criterions for energy and force. We start from various initial structures with all degrees of freedom included to see if they will finally converge to the same structure to ensure that we reach the optimal structures.

The Vienna Ab initio Simulation Package (VASP) is used to perform periodic DFT calculations,1-3 employing the projected augmented-wave (PAW) pseudopotentials and the generalized gradient approximation with exchange-correlation functional by Perdew, Burke, and Ernzerhof (PBE).4-6 The conjugated gradient algorithm is chosen to perform the geometric optimization in VASP.

2. H2O Structure

Before we can perform the geometric optimization of H2O, convergence tests are required firstly with respect to the size of the unit cell as well as the energy cutoff. The test for number of k-points is not required here because all the calculations for such geometric optimization of one molecule in a large box use a 1x1x1 k-point sampling. For both H2O and HCN, all three atoms lie in one plane. In order to do the convergence test, we initially put the O atom at the (0, 0, 0) and two H atoms at (1/L, 0, 0) and (-1/L, 0, 0) according to the experimental results (H-O bond length usually around 1 Å) as shown in Figure 1 (L is the side length of the unit cell).

Figure 1. H2O in a large box (H2O is located at the vertex of the cubic unit cell)

 

Energy cutoffs ranging from 400.00 to 600.00 eV are used and the side length of the cubic unit cell is chosen from 20.0 to 30.0 Å with an increment of 1 Å. The results of the convergence tests are summarized below in Table 1-2 and Figure 2-3.

Table 1. Convergence tests on energy cutoff for H2O

Table 2. Convergence tests on unit cell size for H2O

Figure 2. Energy vs energy cutoff for H2O                  Figure 3. Energy vs unit cell size for for H2O

Based on these results, we chose the energy cutoff of 500.00 eV (tolerance of 0.005 eV) and the side length of 25.0 Å for further calculation of the water molecule.

Now, we firstly vary the bond length between H and O atoms while still keeping them linearly arranged on the y-axis. The O atom remains (0, 0, 0) and the two H atoms are located at (a/L, 0, 0) and (-a/L, 0, 0). Here, L=25.0 Å and a is chosen from 0.5 to 1.3 Å with the interval of 0.1 Å. The results are summarized below in Table 3.

Table 3. Results for H2O with different initial bond lengths

 

From these results, we can see that the energies for the cases with initial bond lengths of 0.5 and 0.6 Å are significantly higher as the distances between atoms in the optimal geometry, which are trapped in a local minimum, are much larger than typical bond lengths as shown in Figure 4. The distances between the O atom and the H atom are 9.790 and 4.809 Å for these two cases, respectively. As two H atoms are initially too close to the O atom, this situation physically represents the enormously compressed bonds, which brings a strong repulsive force pushing the atoms away from each other. The numerical optimization method then takes an initial step that separates the two atoms by a very large distance and they can never go back since the force is much weaker now due to the very large distance. All other cases give us the same optimized structure as shown in Figure 5. The H-O bond length is 0.943 Å and the bond angle qHOH is 180.00°.

    

Figure 4. Separated H2O with initially short distances   Figure 5. Optimized structure for H2O

 

In order to ensure that the structure above is the optimal structure, we modify the initial structure through changing the bond angle that does not make components of the forces vanished by symmetry. So, with the O atom remaining at (0, 0, 0), the two H atoms are at (a/L, b/L, 0) and (-a/L, b/L, 0). L here is the side length of 25.0 Å, and according to the results above, the a is fixed at 0.9 Å while b is chosen from 0.1, 0.2, and 0.3 Å to have three different initial bond angles. After optimization, we find that these three tests finally converge at the same free energy of -14.217 eV with the same optimized structure of H2O as shown in Figure 6. The H-O bond length is 0.972 Å and the bond angle qHOH is 104.50°. The energy is lower than the previous cases with bond angle of 180.00° and since three different bond angles finally converges to the same structure, this is the optimal structure for H2O. The experimental H-O bond length and bond angle is 0.958 Å and 104.48°, respectively. We see that the DFT-predicted bond length and bond angle are very similar to the experimental results with differences less than 2%.

Figure 6. Optimal structure for H2O

 

3. HCN Structure

Similarly, the convergence tests are required before we optimize the structure HCN. Due to the similar size of these two molecules, the convergence test on unit cell sizes is no longer required and the side length of 25.0 Å above is used. We similarly put the C atom at the (0, 0, 0), the N atom and the H atom are at (1/L, 0, 0) and (-1/L, 0, 0) as shown in Figure 7 (L is the side length of the unit cell and is fixed at 25.0 Å here).

Figure 7. HCN in a large box (HCN is located at the vertex of the cubic unit cell)

 

Energy cutoffs ranging from 400.00 to 600.00 eV are tested for HCN. The results are summarized below in Table 4 and Figure 8. Based on these results, we chose the energy cutoff of 500.00 eV using the same tolerance for further calculations of HCN.

Table 4. Convergence tests on energy cutoff for HCN

Figure 8. Energy vs energy cutoff for HCN

 

Again as we did for the H2O molecule, we vary the N-C and H-C bond lengths while keeping all of them on the y-axis. The C atom remains (0, 0, 0) and the N and H atoms are located at (a/L, 0, 0) and (-a/L, 0, 0). Here, L=25.0 Å and a is chosen from 0.7 to 1.5 Å with the interval of 0.1 Å. The results are summarized below in Table 5.

Table 5. Results for H2O with different initial bond lengths

 

From these results, we can see that the energies for the cases with initial bond lengths of 0.7, 0.9, and 1.0 Å are significantly higher due to the large distances between atoms as shown in Figure 9. These atoms are pushed apart from each other with the similar reason for those H2O with very short bond lengths (only one H-C bond, the N atom is far away). For the case of 0.8 Å, it even cannot converge. The optimized structures for the rest cases are totally the same as shown in Figure 10. The N-C and H-C bond lengths are 1.161 and 1.075 Å, respectively. The bond angle qHCN is 180.00°.

             

Figure 9. Separated HCN with initially short distances    Figure 10. Optimized structure for HCN

 

Similarly, we then modify the initial structure by changing the bond angle to make sure that the force components are not canceled out by symmetry. So, with the C atom remaining at (0, 0, 0), the N and H are at (a/L, b/L, 0) and (-a/L, b/L, 0). L here is the side length of 25.0 Å, and according to the results above, the a is fixed at 1.2 Å while b is chosen from 0.1, 0.2, and 0.3 Å to have three different initial bond angles. After optimization, we find that all these three tests finally give the same free energy of -19.690 eV and the same optimized structure of HCN as shown in Figure 11. The N-C and H-C bond lengths are 1.161 and 1.075 Å, respectively. The bond angle qHCN is 179.95°. The energy is almost the same compared to the previous cases with bond angle of 180.00° and since three different bond angles finally converges to the same structure, this is the optimal structure for HCN. The experimental N-C and H-C bond lengths are 1.156 and 1.064 Å, and the bond angle 180.00°. We see that the DFT-predicted bond length and bond angle are very similar to the experimental results with differences less than 2%.

Figure 11. Optimal structure for HCN

 

4. Conclusion

To conclude, in this project, we find the optimal structures of H2O and HCN. For H2O, the DFT-predicted H-O bond length is 0.972 Å and the bond angle qHOH is 104.50° (the experimental values are 0.958 Å and 104.48). For HCN, the DFT-predicted N-C and H-C bond lengths are 1.161 and 1.075 Å while the bond angle qHCN is 179.95° (the experimental values are 1.156 and 1.064 Å, and 180.00°). We see that the DFT can accurately predict the bond lengths and bond angles for these selected molecules with differences less than 2% compared to experimental results.

 

Reference

  1. Kresse, G. and J. Hafner, Ab initio molecular dynamics for liquid metals. Physical Review B, 1993. 47(1): p. 558.
  2. Kresse, G. and J. Hafner, Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Physical Review B, 1994. 49(20): p. 14251.
  3. Kresse, G. and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Physical review B, 1996. 54(16): p. 11169.
  4. Kresse, G. and D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method. Physical Review B, 1999. 59(3): p. 1758.
  5. Blöchl, P.E., Projector augmented-wave method. Physical review B, 1994. 50(24): p. 17953.
  6. Perdew, J.P., K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple. Physical review letters, 1996. 77(18): p. 3865.

Vibrational frequencies of water

NameFunctional (LDA-PWC or GGA-PBE)Bending (cm-1)Stretching Symmetric (cm-1)Stretching Asymmetric (cm-1)
Haoran HeGGA PBE1668.433672.103786.41
Jeff RableGGA PBE1614.683739.133853.83
Mohamed Mauroof UmarGGA PBE1614.723739.003853.69
Brett GreenGGA-PBE1614.723739.003853.69
Bryan BrasileGGA-PBE1631.003686.003806.00
Garrett DuCharmeGGA-PBE1614.503740.103854.90
Jonathan SchirmerGGA-PBE1679.003663.003778.00
Mark BronsonGGA-PBE1609.603713.203823.40
Mark HagemannGGA-PBE1614.803739.103853.80
Ran ChenGGA-PBE1614.803739.103853.80
TongzhouGGA-PBE1617.583743.803857.37
Zihao ChenGGA-PBE1617.023743.523857.61
Haoran HeLDA PWC1631.233685.453805.69
Jeff RableLDA PWC1579.233750.543871.47
Mohamed Mauroof UmarLDA PWC1579.283750.403871.31
Brett GreenLDA-PWC1579.653748.723869.57
Bryan BrasileLDA-PWC1669.003672.003787.00
Garrett DuCharmeLDA-PWC1579.303750.603871.50
Jonathan SchirmerLDA-PWC1634.513675.543795.73
Mark BronsonLDA-PWC1578.803727.203838.50
Mark HagemannLDA-PWC1579.303750.503871.50
Ran ChenLDA-PWC1579.303750.503871.50
TongzhouLDA-PWC1580.053752.843872.85
Zihao ChenLDA-PWC1579.283750.403871.31

Status of Posts 2 due on March 1st

ReviewerAuthorPost Ready for Review (Due 2/20)Review Done (Due 2/27)Post Published (Due 3/1)
Brasile, Bryan WilliamXiao, Run2018-02-25 (Late excused - Comprehensive exam)2018-02-262018-03-02
Bronson Jr., Mark JosephZhao, Tongzhou2018-02-202018-02-272018-03-02
Chen, RanBrasile, Bryan William2018-02-282018-03-012018-03-02
Chen, SiBronson Jr., Mark Joseph2018-02-202018-02-252018-03-01
Chen, ZihaoChen, Ran2018-02-222018-02-242018-04-03
Ducharme, Garrett LouisChen, Si2018-02-212018-02-262018-02-27
Green, Brett RobertChen, Zihao2018-02-202018-02-232018-02-25
Hagemann, Mark AlexanderDucharme, Garrett Louis2018-02-212018-02-222018-02-01
He, HaoranGreen, Brett Robert2018-02-222018-02-222018-02-23
Rable, Jeffrey GHagemann, Mark Alexander2018-02-222018-02-232018-02-27
Schirmer, Jonathan ProssickHe, Haoran2018-02-212018-03-012018-03-01
Umar, MohamedRable, Jeffrey G2018-02-202018-03-012018-03-01
Xiao, RunSchirmer, Jonathan Prossick2018-02-202018-02-262018-04-20
Zhao, TongzhouUmar, Mohamed2018-03-022018-03-052018-03-06

Determination of Pt crystal structure and corresponding lattice constant

In this project, the energetically favorable crystal structure of Pt and corresponding lattice constant were determined using Density Functional Theory. The total energies were computed in Material Studio with CASTEP Calculation Package [1]. The functional of Perdew Burke, and Ernzerhoff was employed [2]. A plane wave basis set was used with a cut-off energy of 321.1 eV and OTFG ultrasoft pseudopotentials was solved using Koelling-Harmon treatment. The pseudo atomic calculation was performed for Pt 4f14 5s2 5p6 5d9 6s1.

Determine Energy Cut-off
The energy cut-off was determined by considering both calculation accuracy and computational cost. Pt fcc with lattice constant of 3.21 angstrom was randomly selected to perform a series of calculation, in order to determine energy cut-off. As shown in Fig.1, the fluctuation of cohesive energy becomes smaller as energy cut-off increases, while the computational cost has an increasing trend. In order to guarantee the accuracy and stability of our data and also keep computational cost acceptable, we choose energy cut-off 321.1 eV.


Fig.1 a.cohesive energy for 12 energy cut-off values b. computational time for 12 energy cut-off values

Determine K Points
Similar to energy cut-off determination, determining the number of K points was also based on calculation accuracy and calculational expense. From Fig.2, it is clear that computational cost increase with number of K points and cohesive energy becomes more stable with the energy difference of 0.05 eV between 20 and 28 kpoints


Fig.2 a.cohesive energy for 6 K points values b.Computational time for 6 K points values

Determine Pt Crystal Structure and Lattice Constant
Simple cubic structure
For simple cubic structure, 10 × 10 × 10 K points were sampled with 0.1 eV Gaussian smearing width. The energetically most favorable lattice constant is 2.6 Å with cohesive energy of -9.175 eV.

Fig.3 Cohesive energies of Pt in simple cubic structure as a function of lattice parameters

Face center cubic structure
For face center cubic structure, 12 × 12 × 12 K points were sampled with 0.1 eV Gaussian smearing width. The energetically most favorable lattice constant is 3.924 Å with cohesive energy of -9.667 eV.

Fig.4 Cohesive energies of Pt in fcc structure as a function of lattice parameters

Hexagonal Close Packed (hcp) Structure
When performing calculations of Pt hcp, two parameters had to be considered. We randomly selected three lattice constants of a and 12 c values for corresponding a. 12 × 12 × 8 K points were sampled with 0.1 eV Gaussian smearing width. The energetically most favorable lattice constant is 2.7 Å and height of 5.4 Å with cohesive energy of -9.512 eV.

Fig.5 Cohesive energies of Pt in hcp structure as a function of lattice parameters

Conclusion
From calculations performed above, we can conclude fcc structure with lattice constant of 3.924 Å is energetically most favorable for Pt, which is in a good agreement with experiment 3.912 Å [3].

Reference
[1] “First principle methods using CASTEP” Zeitschrift fuer Kristallographie 220(5-6) pp. 567-570 (2005)
[2] Perdew, J. P; Burke, K; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865-3868
[3]”Precision Measurement of the Lattice Constants of Twelve Common Metals” Davey, Wheeler, Physical Review. 25 753-761 (1925)

 

 

Determining the lattice constants for Hf

   For Project 1,  the lattice constants for Hf are calculated. Since Hf is an hcp metal with c/a=1.58, we need to calculate the energy by changing the lattice constant c while keeping c/a fixed. And the optimized a, c can be obtained by minimizing the free energy.

1. The hcp structure of Hf

   Figure 1 shows the unit cell of Hf hcp structure. OA, OB, OC are three lattice vectors. The lengths and angles of the lattices vectors (OA, OB, OC) are a, b, c and α, β, γ, respectively. To keep the symmetry, the constraints for the unit cell are a=b, α=β=90 degrees, γ=120 degrees.

Figure 1 The unit cell of Hf

2. The parameters for the calculations

Some important parameters and inputs for the calculation are listed as following:

Atomic structure for Hf:1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 4f14 5s2 5p6 5d2 6s2

The core configuration of the pseudopotential: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10

Functional: GGA Perdew Burke Ernzerhof (PBA) functional

Pseudopotential: OTFG ultrasoft

Smearing width: 0.1

Origin shift of K points: 0, 0, 0

The k points grid: investigated in the next section

The energy cutoff: investigated in the next section

3. The convergence of the calculations

   Before starting the actual calculation for the lattice constant, it is important to explore the convergence of the calculations with respect to the number of k points and the energy cutoff. So the most accurate and efficient results can be obtained. For the following calculation, the lattice constant is fixed at c= 5.0511Å, a= 3.1946Å, while the size of k point grid and the energy cutoff are varied.

   (1) The number of k-points in the irreducible part of the BZ

     In this section 3.1, the energy cutoff is fixed at 400 eV, and the size of k point grid changes and Number of k-points in the irreducible part of the BZ (the number of k points) varies from 1 to 162, shown in Table 1. According to Table 1, the relations between Energy/atom, calculation time and the number of k points can be plotted, which are shown in Figure 2. As shown in Figure 2, the energy/atom becomes stable with the increase of k points. However, the calculation time also increases rapidly when the number of k points increases. By combining the results of these two graphs, 9*9*6 k points with 36 k points is chosen to obtain an accurate result with less calculation time.

   A little discussion here, since my calculation involves unit cells with different volumes, the k points need to be chosen so the density of k points in reciprocal space is comparable for the different supercells. But as shown in section 4, since the volumes of unit cell change just a little during the calculation, fixing the number of k points will not make much difference. So in section 4, the number of k points is chosen as 36 for all calculations.

Table 1 The calculation of different sizes of k point grid

Figure 2 The plots between Energy/atom, calculation time and the number of k points. The solid lines are just used for guide the eye.

(2) The energy cutoff

In this section, the number of k point is fixed at 36, and the energy cutoff varies from 200 to 800, shown in Table 2. And Figure 3 shows the relation between energy/atom, calculation time and energy cutoff. Although the energy/atom keeps increasing with the increase of energy cut off, the differences between each energy/atom are becoming smaller and smaller. On the other hand, the calculation time increases rapidly with the increase of energy cut off. In order to save the time and keep the accuracy, the energy cutoff is chosen as 435.4 eV.

Table 2 The calculation of different energy cutoff

Figure 3 The plots between energy/atom, calculation time and energy cutoff. The solid lines are used to guide the eye.

4 The calculation of the lattice constants

Using the results from section 3, the k points and the energy cutoff are chosen as 9*9*6 and 435.4 eV. Then lattice constants c and a are changed to obtain the minimum energy, shown in Table 3. As shown in Figure 4, the unit cell and the lattice constants can be predicted. As shown in Figure 4a, the quadratic relation between energy/atom and volume/atom only valid for a small range of lattice constants around the equilibrium value. A small range of volume/atom is chosen and the relation is re-plotted in Figure 4b. According to the fitting equation in Figure 4b, the fitting volume/atom is 22.64 Å^3 and the lattice constant are c=5.0673 Å, a=3.2072 Å. And for the experiment data, the lattice constants are a=3.1964 Å, c=5.0511 Å (1). So the calculation is accurate with an error of 0.3%.

Table 3 The calculation of energy with different lattice constants.

Figure 4 The plots between energy/atom and volume/atom. the solid lines are used to guide the eye. The dashed line is the fitting line.

5. Some discussion and future works

Although the calculation result is accurate with an error of 0.3%, there are still some ways to further increase the accuracy. First, the energy cut off can be increased so the energy/atom can be more precise. Second, we can calculate more points around the value c=5.05 Å and maybe a more accurate result can be obtained. Third, instead of using the quadratic fitting, we can use a higher order polynomial fitting.

For some future works: first, it is worthwhile changing the ratio c/a to determine the optimum value of c/a. Second, it should be interesting to analyze the behavior of energy at small or large lattice constants.

6. Reference:

(1) https://www.webelements.com/hafnium/crystal_structure.html

Determination of Structure and Lattice Constant for Platinum

Overview of the Problem

This post will be about determining both the structure and lattice constant for platinum. This will be done using density functional theory (DFT) to calculate the cohesive energies across a range of lattice parameters for platinum in the simple cubic (SC), face-centered cubic (FCC), and hexagonal close-packed (HCP) structures.

In order to make sure the results are feasible, convergence tests are first done for both the number of k-points and the cutoff energies.

All calculations are done in Materials Studio with the CASTEP DFT package [1]. The functional used was that of Perdew Burke Ernzerhof [2], and the pseudopotential was obtained using the Koelling-Harmon solver for the 4f14 5s2 5p6 5d9 6s1 outer shells. 

Cutoff Energy Convergence

The cutoff energy was tested by starting at a cutoff energy of 360 eV and then increasing by increments of 30 eV to look for convergence in the free energy. For this test the lattice parameters of the three structures were chosen as a = 3.92 Å for FCC, a = 2.62 Å for SC, and a = 3.02 Å, c = 4.83 Å for HCP. These lattice constants were chosen for this test by roughly estimating the location of the minimum of the three structures at an energy cutoff of 300 eV. The results are outlined in table 1 and figure 1.

 

Energy Cutoff (eV)FCC Free Energy (eV)SC Free Energy (eV)HCP Free Energy (eV)
300-13050.818-13050.421-13049.923
330-13050.927-13050.534-13050.046
360-13050.960-13050.568-13050.080
390-13050.970-13050.578-13050.116
420-13050.973-13050.580-13050.119
450-13050.973-13050.581-13050.120
480-13050.974-13050.581-13050.120
510-13050.974-13050.581-13050.120
Table 1: Data for the energy cutoff convergence test.

As can be seen from the figure and the table, a lower energy cutoff here will overestimate the free energy. It can also be see that the variation in the free energy is less than 0.01 eV past the 390 eV cutoff energy. For consistency, a cutoff energy of 420 eV has been chosen for all calculations involving the determination of the lattice parameters.

K-Point Convergence

The k-point convergence was tested using an HCP structure with a = 3.12 Å and c/a = 1.6, an SC structure with a = 2.62 Å, and an FCC structure with a = 3.92 Å. The results are outlined in tables 2, 3, and 4, as well as figures 2, 3, and 4.

K-point GridNumber of K-pointsFree Energy (eV)
7x7x432-13050.000
8x8x5156-13049.970
9x9x675-13049.970
10x10x6240-13049.965
11x11x7144-13049.955
12x12x8456-13049.975
13x13x9196-13049.970
Table 2: K-point convergence data for HCP.
K-point GridNumber of K-pointsFree Energy (eV)
6x6x628-13050.960
8x8x820-13050.94
10x1010110-13050.943
11x11x1156-13050.945
12x12x12182-13050.945
13x13x1384-13050.943
14x14x14280-13050.943
15x15x15120-13050.945
16x16x16408-13050.945
Table 3: K-point convergence data for FCC.
K-point GridNumber of K-pointsFree Energy (eV)
10x10x1035-13050.460
11x11x1156-13050.470
12x12x1256-13050.490
13x13x1384-13050.500
14x14x1484-13050.510
15x15x15120-13050.500
16x16x16120-13050.500
Table 4: K-point convergence data for SC.

Figure 2: K-point convergence plot for SC

Figure 3: K-point convergence plot for HCP

Figure 4: K-point convergence plot for FCC

 

For the SC lattice, having few k-points gives energies that start above the convergence level with a tendency to decrease. This is not the case for the HCP lattice, where the energies start low and then increase before decreasing again. Lastly, for the FCC lattice we see that the energy starts above the convergence level before sharply decreasing and then approaching convergence.

Based on the results from the test of the k-point convergence, grids of 15x15x15 were chosen FCC and SC, while a grid of 13x13x9 was chosen for HCP.

Lattice Parameter for SC

The simple cubic structure is shown in figure 5.

 

Figure 5: Simple cubic lattice structure

The calculations for the cohesive energy were done by first calculating the free energy per atom for the simple cubic structure, and then subtracting away the atomic energy of -13042.12 eV, which was obtained from the pseudopotential calculations. The results are shown in figure 6.

 

Figure 6: Cohesive energy vs lattice parameter for the simple cubic structure.

Pt in the simple cubic configuration is found to be stable with a lattice constant of a = 2.62 Å and a cohesive energy of -8.396 eV.

Lattice Parameter for FCC

The face-centered cubic structure is shown in figure 7.

Figure 7: The face-centered cubic lattice structure

 

The calculations of the cohesive energy were done in the same way as they were for the simple cubic lattice. The atomic energy will remain the same, so we simply calculate the free energy and take the difference. The results of these calculations are shown in figure 8.

Figure 8: Results from the calculation of cohesive energy for Pt in the FCC configuration for various lattice parameters

From the results of the calculation, it is found that Pt in the FCC configuration has a preferred lattice constant of about 3.96 Å, which gives a cohesive energy of -8.851 eV.

Lattice Parameter for HCP

The lattice of the hexagonal close-packed structure is shown in figure 9.

 

Figure 9: The hexagonal close-packed lattice structure

The HCP lattice has two lattice constants, so there is a much larger phase space to explore in order to locate the minimum cohesive energy. In order to sample this space, the ratio between the lattice constants, c/a, is held fixed at values of 1.57, 1.6, and 1.63. The parameters a and c are then varied in tandem to search for the preferred lattice constant. The results of these calculations are shown in figure 10.

Calculations for the cohesive energy of HCP Pt for c/a ratios of 1.57, 1.6, 1.63.

The results show that an HCP lattice of Pt is most stable with lattice constants of a = 2.98 Å and c = 4.77 Å for c/a = 1.6. These parameters give a cohesive energy of -7.996 eV.

Conclusions

The calculations done above give cohesive energies of -8.396 eV for SC,  -8.851 eV for FCC, and -7.996 for HCP. This implies that Pt is most stable in the FCC configuration with a lattice constant of 3.96 Å. Davey finds through experimental methods that Pt is most stable in the FCC configuration with a lattice constant of 3.91Å [3]. Using density functional theory, we have shown that we can correctly predicted that Pt naturally forms an FCC lattice. However, we have overestimated the lattice constant by about 0.05 Å.

 

References

[1] First principle methods using CASTEP Zeitschrift fuer Kristallographie 220(5-6) pp. 567-570 (2005).

[2] Perdew, J.P., K. Burke, and M. Ernzerhof, Generalized gradient approximation made simple.Physical review letters, 1996. 77(18): p. 3865.

[3] Davey, W. P. (1925). Precision Measurements of the Lattice Constants of Twelve Common Metals. Physical Review, 25(6), 753-761. doi:10.1103/physrev.25.753.

Determination of Lattice Constant of Platinum Cell

Introduction

Platinum is a chemical element with symbol Pt and atmomic number 78. The electron configuration is [Xe] 4f145d96s1. Its crystal structure is face-centered cubic (fcc, see Fig. 1), and the lattice constant is 3.9239 Å[1].

The target for this computation experiment is to use density functional theory (DFT) to evaluate the energy of fcc lattice at different lattice constants. The estimation of the lattice constant is hence decided by minimizing the energy. The result is compared with the standard data based on experimental measurements.

Key Factors of Density Functional Theory (DFT)

Denstity functional theory (DFT) is a computational method based on calculating the energy of a condensed matter system as a functional of the electron density. Thereare a lot of references that give systematic introductions and discussions on the method, which we will not go into details in this report[2]. The key factors affecting the calculation in this report include (1) Energy cut-off and (2) k-point sampling number. Energy cut-off refers the maximum energy of the bands involved in our calculation. The number of the bands included for a typical calculation is about 20, and this number may increase when higher energy resolution is required. K-point sampling number refers the number of sampling points within the first Brillouin zone. Usually the higher the energy cut-off is and the larger the sampling number is, the better the accuracy is. Due to limitation on the computational resources, we have to decide lowest values of both the factors that could meet requirements.

We performed several trial calculations and compared their numerical variance to decide the appropriate values. (1) We fix the cut-off energy at 800 eV, which is high enough for a simple calculation, and explore the effect of k-point density on energy calculation and atomic energy to figure out the minimum sufficient value for k-point sampling number. (2) We use the number of k-point sampling mentioned above and explore the effect of cut-off energy and decide the best cut-off energy that balances the calculating speed and precision. Both tests are performed at conventional lattice constant a=4Å.

Energy Evaluation and Lattice

Applying the parameters decided above, several calculations of free energy at lattice constant ranging from 3.4 ~ 4.4Å were performed. Then a quadratic fitting was performed to minimizing energy. The corresponding lattice constant is our theoretical prediction of the true lattice, and it is compared with standard value. A more detailed discussion can be found in the experimental section in this report.

Experimental Facts

The basic parameters and key factors shared in all the calculations are listed below:

Lattice cell type: primitive cell of face centered cubic (fcc)

Functional Type: GGA PBE

Smearing width: 0.10

Pseudo Potential Orbitals: 4f14 5s2 5p6 5d9 6s1

Pseudo Type: OTFG ultrasoft

k-points origin shift: no

K-point Number Examination

Based on the discussion above, we calculated the free energy of the lattice at a=4Å. The results are shown in Fig. 2. Based on this result, it is sufficient to choose 16×16×16 as the appropriate sampling number for one wants acceptable precision without spending too much time.

Energy Cut-off Examination

The convergence of energy and pseudo atomic energy are displayed as follows. The difference between them is the binding energy per atom. Although it is not fully converged, the pseudo atomic energy variance is highly suppressed when the cut-off energy is higher than 800eV, which indicates that 800eV is a reasonable choice for our estimation of the binding energy (Fig.3 and Fig.4).

 

 Energy at Different Lattice Constants

To best fit the plots in the graph, a polynomial to the order of 3 was applied, and the minimum free energy was reached at lattice constant a=3.95Å. Compared to the standard result, this calculation has an error of 0.6%.

Reference

[1] https://en.wikipedia.org/wiki/Platinum

[2]Sholl, David, and Janice A. Steckel. Density functional theory: a practical introduction. John Wiley & Sons, 2011.