Category Archives: HCN and H2O

Determination of the Geometry of Water Molecule

Project Description

Water molecule is made of 1 oxygen atom and 2 hydrogen atoms. The oxygen is covalently bonded to the two hydrogens. In this project, we apply density functional theory (DFT) method to optimizing the geometry of water molecule and find the correct distance between hydrogen atoms and oxygen atom as well as the angle between two bonds. The is done by doing an self-consistent calculation which minimizing the energy of the system.

Method

To correctly imitate the behavior of molecules, one essential step of our calculation is to design a super lattice that work as the container of the isolated molecules, which allow us to perform plane wave calculation and apply periodic boundary condition [1]. This is done by creating a crystal lattice and attach the oxygen atoms and hydrogen atoms on the lattice. One must avoid apply any artificial symmetry onto the system, because in nature all the symmetric properties of the molecule are the results of lowering the energy. Based on this principle, we choose triclinic lattice. To avoid the influence of adjacent molecules, we set the the size of the lattice far bigger than the size of the molecules. Hence we attach oxygen atom at fractional position at  (0,0,0) and hydrogen atoms at (0, 0.1, 0) and (0.1, 0, 0). The fractional coordinates of all the atoms are fixed during the energy optimization. The fraction is chosen such that the distance between neighboring molecules is roughly 10 times bigger than the size of the molecule, which is a good imitation of the gas in nature. In principle, this is all we need to start the calculation. The software will adjust the lattice size and the angles such that the structure of the lattice is energetically favorable. In our calculation, we also fix the interlayer distance c to a very big value and α, β to 90° (Fig 1). Since the only effect of these parameters is energy between different molecules, we believe this will have little effect on our result as long as c is big enough.

Fig 1 Initial settings of the lattice cell. The distance between two layers is c=10Å. The only angle that is allowed to change is the bond angle between two hydrogen atoms

Numerical Calculation and Results

We use Materials Studio® to perform the calculation. The parameters that are not mentioned are all set by default.

The following are the detailed steps and parameters we applied in our calculation:

Lattice type and symmetry group: Triclinic, 1 P1, with a, b, c all set to 10Å, α, β, γ all set to 90°.

Fractional positions and constraints: O (0, 0, 0), H (0.1, 0, 0), H (0, 0.1, 0). Interlayer distance c and angle α, β are fixed, fractional positions are fixed during the optimization.

Functionals: CASTEP calculation with functional GGA PBE.

Pseudo atomic orbitals: O  (2s2 2p4),  H (1s1)

Geometry optimizing method: BFGS, full cell optimization.

k-point griding: 1×1×1

Energy cutoff: up to 900eV

The most important factor that could affect the result is the energy cutoff. Since a large amount of the lattice is vacuum, more plane waves are required to obtain a relative accurate result. We examine the result by performing several calculations at different cutoffs. The number of k-points in our calculation is not important, because of the large size of the unit cell. The results are shown in the table. 1 and figure. 2. From the calculation we know that energy cutoff 800 eV is high enough to obtain a bond angle up to 0.01Å and bond bond angle up to 0.1°. The experimental result for water molecule is that the bond length is 0.96Å and the bond angle is 104.5°. The relative error for bond length and bond angle are 1% and 2% respectively. In addition, our calculation does not assume any symmetry of the molecule, yet the calculation gives the same bond length for both hydrogen atoms, this shows that the symmetric structure is energetic favorable and it agrees with the experiment.

Table 1 Geometry optimization of water molecule at different energy cutoffs

Energy cutoff /eVBond length a /ABond length b /ABond angle /Degree
571.40.970.97102.2
6000.970.97102.0
7000.970.97102.1
8000.970.97102.2
9000.970.97102.1

Fig 2 Lattice cell after energy optimization. After the geometry optimization, the shape of the lattice is distorted. The bond angle is 102.2°, and the bond length is 0.97Å.

We also compared the result with that obtained from traditional method. In traditional method, not the lattice but the atom positions are movable during the geometry optimization, which we call as ‘fixed lattice method’. For simple molecules such as water, the two methods give agreeable results (Fig. 3). However, it is recommended to use the fixed lattice method because (1) when the molecule is made of multiple atoms there would be multiple positions to optimize, and (2) the speed of lattice optimization is slower than the usual method (more than 600s comparing to 200s in the usual method).

Fig.3 Fixed lattice geometry optimization. The lattice itself is fixed while the atoms’ positions are optimized. The bond length converged to 0.97Å and bond angle is 104.2°.

 

Conclusion

In our report, we introduce a method to optimize the geometry of water molecule by adjusting the lattice. Our result shows the method can give correct results. We also compare the this method with the usual method which fixes the lattice, and we find the usual method has a better performance (less time) and better portability (applies for more complicated molecules). Although the method implanted in this report is not so successful, one valuable conclusion we can get is that when we do DFT calculation, it is better to have as many as possible constraints on lattice and symmetry such that the variance on the density functionals is minimal.

Reference

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

[2] https://en.wikipedia.org/wiki/Properties_of_water#Polarity_and_hydrogen_bonding

Determining the optimized geometry of water molecule

1. Introduction of the project

In this project, I calculated the optimized geometry of H2O. To get it, first, a cubic supercell (side length L angstroms) that is mostly empty space is built. Then the oxygen atom is placed in the supercell with fractional coordinates (0,0,0) and two hydrogen atoms are placed in the super cell with fractional coordinates (a/L, b/L,0) and (-a/L, b/L,0), shown in Figure 1. Next, by calculating the geometry optimization using CASTEP [1], the optimized geometry can be obtained, which includes the bond length and the bond angle.

Figure 1 The initial supercell structure. The red atoms are the oxygen atoms and the grey ones are hydrogen atoms.

During the calculation, the energy cutoff, the k point grid, the supercell lattice L can be changed to obtain most efficient and accurate results. The initial fractional coordinates of H can be changed to analyze if multiple initial geometries for each molecule converge to the same final state.

2. The parameters for the calculations

Some important parameters and inputs for the calculation are listed as following:
Atomic and pseudo atomic structure for H: 1s1
Atomic structure for O: 1s2 2s2 2p4
pseudo atomic structure for O: 2s2 2p4
Functional: GGA Perdew Burke Ernzerhof (PBA) functional
Pseudopotential: OTFG ultrasoft
The k points grid: investigated in the section 3
The energy cutoff: investigated in the section 4

3. The energy cut off

We can start with the k point grid of 3*3*3 and the supercell lattice L=10 Å. The initial fractional coordinates of H are chosen as (0.1, 0.05, 0) and (-0.1, 0.05, 0). From the calculation results shown in table 1, the energy cutoff can be chosen as 700eV.

Table 1 The calculation of the energy cut off.

4. The k point grid

We start with the energy cut off =700eV and the supercell lattice L=10 Å. The initial fractional coordinates of H are chosen as (0.1, 0.05, 0) and (-0.1, 0.05, 0). From the calculation results shown in table 2, the k point grid can be chosen as 1*1*1. It is easy to understand because the supercell is too large and we only consider the isolated molecule instead of the periodic system.

Table 2 The calculation results of the k point grid.

5. The supercell length L

Since we want to get the optimized geometry for an isolated molecule, the supercell lattice L need to be large enough to separate neighbouring molecules. As shown in the Table 3, the supercell length is varied, and the final supercell lengths, final bond length, bond angle are also changed. When initial supercell length is bigger than 10 Å, the final supercell lengths, final bond length and bond angle will become stable, which means the molecules are separated and they cannot feel the forces from neighbouring molecules. So we can choose the supercell length as 10 Å.

Table 3 The calculation results of the supercell length.

6. The initial hydrogen atom positions

The initial and final bond angle and bond length are changed as shown in Table     . As seen in table if we choose the initial bond angle equals 180°, the final bond angle will also be 180° . Because in this situation, the force along atoms direction is already balanced so the angle will not change. But if the initial bond angle is not 180°, the final bond length and angle are always 0.970 Å and around 104.5°.

Table 4 The calculation results for different initial hydrogen atom positions.

7. Conclusion

As shown in Figure 2, under the condition that the energy cut off is 700eV, the k points grid is 1*1*1, the supercell lattice length is 10Å, the initial bond length is 1.118 Å and the initial bond angle is 126.9°, the final results for bond length and angle are 0.970 Å and  104.582°. The experimental results for water molecule are: bond length is 0.958 Å bond angle is 104.45°. So the error for the bond length is about 1.2% and the error for the bond angle is about 0.1%.

Figure 2 The final water molecule structure for initial bond length is 1.118 Å and initial bond angle is 126.9°.

8. Discussion and future works

Although the results are very accurate with small errors, there are still some ways to further increase the accuracy. For example, during the calculation of geometry optimization, the parameters of convergence tolerance, such as Max. force, Max displacement, can be changed to increase the accuracy.

For some future works: Hydrogen bonding exists between water molecules, which will affect the water molecule structure, so it is worthwhile to try to calculate the Hydrogen bonding. Also since the water molecule has the dipole momentum, we can also calculate the dipole momentum of water molecules and compare it with the experimental results.

Reference

[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] The water molecule structure. https://www.worldofmolecules.com/solvents/water.htm

Exploring Geometry Optimization with a Water Molecule

Introduction

The focus of this post will be to explore different aspects of geometry optimization by studying a water molecule. The most stable geometry of water has an O-H bond length of 0.957 Å and an H-O-H bond angle of 104.2° [1]. We will use this fact to test various parameters of the simulation. The geometry optimization calculations were done with the CASTEP density functional theory (DFT) package in Materials Studio [2]. The CASTEP plane wave basis set requires periodic boundary conditions, which creates issues when studying a single molecule. One must carefully consider the size of the simulation cell being used so that the periodic images of the molecule, as seen in fig. 1, do not interact.

Fig. 1: A single water molecule and some of its periodic images.

All calculations use the GGA with the PBE functional [3]. Atomic and pseudo atomic calculations for hydrogen both use the 1s1 electron. Atomic calculations for oxygen are done for the 1s2 2s2 2p4 electrons, while the pseudo atomic calculations are done for the 2s2, 2p4 electrons. The SCF tolerance is 5.0*10^-7 eV/atom and the convergence tolerance is set to 5.0*10^-6 for energy, 0.01 eV/Å for the max force, 0.02 GPa for the max stress, and 5.0*10^-4 Å for the maximum displacement. Unless otherwise specified, the initial conditions are chosen so that the oxygen atom is centered at the origin, while the two hydrogen atoms are at positions (0, 0.707 Å, 0.707 Å) and (0, 0.707 Å, -0.707 Å), giving an initial bond length of 1 Å and an H-O-H angle of 90°.

The rest of the post will be organized into sections focused on understanding the effects of the unit cell size, the energy cutoff, the k-space sampling, variations of the initial conditions, and a discussion of the final results.

Unit Cell Size

The calculations are done with a cubic unit cell. In order to understand the effects that the images have on each other, we vary the lengths of the cube sides from 7.5 Å to 14 Å in increments of 0.5 Å. Each one of these calculations is done with just the gamma k-point and an energy cutoff of 400 eV.

Fig. 2: Variation of the H-O-H bond angle with respect to the lattice parameter.

By looking at the H-O-H bond angle, we can get an idea of how big the unit cell needs to be in order to minimize the role that the images play in the geometry optimization. Fig. 2 shows that this angle changes rapidly at first, but then starts to stabilize at around 12 Å. A table of the data seen in this plot can be found in the appendix. Unit cells with lattice parameters larger than 12 Å take longer to converge, so this is the lattice parameter that will be chosen for the rest of the calculations.

In order to further confirm that this is an appropriate unit cell size, we can also check to see if the energy has converged.

Fig. 3: Energy of the unoptimized water molecule vs the lattice parameter.

The energy changes by less than 0.1 eV once the lattice parameter is greater than 9 Å. Exact values for these calculations are provided in the appendix.

Energy Cutoff

The calculations for testing the energy cutoff were again done with a k-space sampling that only includes the gamma point. We observe changes in the H-O-H bond angle in increments of 20 eV from 400 eV to 660 eV. A table of these calculations can be found in the appendix.

Fig. 4: Variation of the H-O-H bond angle with respect to the energy cutoff.

As can be seen in fig. 3, the H-O-H bond angle does not have any significant change past an energy cutoff of 600 eV.

We can once again further verify that this is an acceptable energy cutoff by looking at the energy of the unoptimized water molecule as we increase the energy cutoff.

Fig. 5: Energy of the unoptimized water molecule vs the energy cutoff.

The energy changes by less than 0.1 eV between an energy cutoff of 560 eV and 600 eV. Exact values for the energies can be found in the appendix.

K-space Sampling

Due to the high symmetry and size of the unit cell, one should expect that the k-space sampling will have little to no impact on the final geometry of the water molecule centered at the origin in real space. We observe that there is no change in the H-O-H bond angle when using 1, 4, or 14 k-points for energy cutoffs of 400 eV as well as 600 eV.

Testing Initial Conditions

Different initial conditions, i.e the initial bond lengths and the bond angles, provide insight into how the geometry optimization is performed. Given a sufficiently large box size and energy cutoff, one would expect that the water molecule should always converge to the optimal geometry regardless of the initial conditions because it has only one stable configuration. For more complicated molecules, it is important to consider the possibility that there are other stable isomers.

Table 1: Effects of initial conditions on the final geometry of the water molecule and the number of optimization steps. All calculations are run with a lattice parameter of 12 Å and an energy cutoff of 600 eV with only the gamma k-point.

The results in table 1 show that starting with an initial angle of 180° will only cause a change in the two bond lengths. This is because the forces between the atoms only need to be minimized along a line instead of within a plane. Furthermore, we see that starting with two different bond lengths will cause the optimization procedure to take longer to achieve the same result. One can see that achieving what appears to be the most physical result in the shortest amount of steps requires the two initial bond lengths to be the same, while the initial angle must be less than 180°.

Discussion

In order to have accurate geometry optimization calculations of molecules with a plane wave basis set, the chosen unit cell must be sufficiently large so that images of the molecule do not interact. Furthermore, it is important that the cell not be arbitrarily large such that the computation time becomes unreasonable. Once this parameter has been tested, investigations into a suitable energy cutoff can be performed. Convergence can be tested yet again by varying the number of k-points. For a small molecule such as water, just the gamma point was shown to be sufficient. Finally, by utilizing symmetries in the system it is possible to decrease the number of optimization steps. However, one must be careful to make sure that the final results are physical, and to also explore the possibility of other stable isomers.

References

  1.  Hoy, A. R. & Bunker, P. R. A precise solution of the rotation bending Schrödinger equation     for a triatomic molecule with application to the water molecule. Journal of Molecular                 Spectroscopy 74, 1–8 (1979).
  2.  Clark, S. et al. First principles methods using CASTEP. Z. Kristallogr. 220, 567–570 (2005).
  3.  Perdew, J. P; Burke, K; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865-3868.

Appendix

Table A1: H-O-H bond angle as a function of the lattice parameter.

Table A2: H-O-H bond angle as a function of the energy cutoff.

Table A4: Energy of the unoptimized water molecule with respect to the energy cutoff.

Table A3: Energy for the unoptimized water molecule with respect to the lattice parameter.

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

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.