Author Archives: Mohamed Mauroof Fadhi Umar

Ab initio calculations to determine the Li – LiH phase diagram

 Theory

In this post we report our findings on our study to determine whether pure Li or LiH is more energetically favorable for different hydrogen pressures P\(_{H_2}\) and temperatures T. For this we need to determine which material minimizes the free energy of the a system containing gaseous H\(_2\) and Li at a given pressure P\(_{H_2}\) and temperature T. We use the grand potential of a given system as a proxy for the free energy since the grand potential of a system is lowest for a system with lowest free energy.

The grand potential for a system containing N\(_{Li}\) Li atoms and N\(_H\) hydrogen atoms is defined by,

\begin{equation}\Omega\left(T,\mu_{Li},\mu_H,N_{Li},N_H\right)=E\left(N_{Li},N_H\right)-TS-\mu_H N_H- \mu_{Li}N_{Li}.\label{gp_def}\end{equation}

Where \(E\left(N_{Li},N_H\right)\) is the internal energy of the material and is determined using a DFT calculation, \(S\) is the material entropy and respectively \(\mu_{species}\) and \(N_{species}\) define the chemical potential and number of atoms of the given species of atoms in the material.

Now we are interested in comparing the grand potentials for Li and LiH. Before proceeding with our calculations we make a few adjustments to equation \eqref{gp_def} to simplify our calculations. First the difference in internal energy between Li and LiH is much larger than the entropic differences between, hence the second term in the right hand side of equation \eqref{gp_def} can be neglected. Next we can eliminate the last term in equation \eqref{gp_def} by choosing to look at  a pure Li unit cell and  a LiH unit cell with equal number of Li atoms. Since \(\mu_{Li}\) is independent of P\(_{H_2}\) and we set \(\left(N_{Li}\right)_{pure\;Li}= \left(N_{Li}\right)_{LiH}\), the term \(\mu_{Li}N_{Li}\) is the same for both systems and will cancel out when considering the difference in grand potential between the two systems.

With above mentioned simplifications equation \eqref{gp_def} now reads as,

\begin{equation}\Omega\left(T,\mu_H,N_H\right)=E\left(N_H\right)-\mu_H N_H.\label{gp_red}\end{equation}

In equilibrium condition the chemical potential of molecular hydrogen and atomic hydrogen are related by,

\begin{equation}\mu_{H}=\frac{1}{2}\mu_{H_2}.\label{muH2_H}\end{equation}.

Treating H\(_2\) as an ideal gas the definition of it’s chemical potential is,

\begin{equation}\mu_{H_2}=E_{H_2}^{total}+\tilde{\mu}_{H_2}\left(T,p^0\right)+k_BTln\left(\frac{p}{p^0}\right).\label{H2_ideal_chem_pot}\end{equation}

Here \(k_B\) is the Boltzmann constant, \(E_{H_2}^{total}\) is the total energy of an isolated H\(_2\) molecule and can be obtained from a DFT calculation for an isolated H\(_2\) molecule. The second term in equation \eqref{H2_ideal_chem_pot}, \(\tilde{\mu}_{H_2}\left(T,p^0\right)\) is the difference between the \(T=0\;K\) and the temperature of interest chemical potential of molecular H\(_2\). \(\tilde{\mu}_{H_2}\left(T,p^0\right)\) can be calculated using data tabulated in NIST-JANAF Thermochemical Tables for molecular hydrogen. In terms of quantities defined in the tables referred to; by noting the chemical potential of an ideal gas is equal to the free energy \(G=H-TS\) of an ideal gas,

\begin{equation}\tilde{\mu}_{H_2}\left(T,p^0\right)=\left[H^0\left(T\right)-H^0\left(T_r\right)\right]-TS\left(T\right)-\left[H^0\left(0\right)-H^0\left(T_r\right)\right].\label{mu_tilde}\end{equation}

Calculations

Now we have the recipe to obtain the Li-LiH phase diagram. From DFT we need the internal energies of a pure Li unit cell, LiH unit cell and a Hydrogen molecule. All energy calculations were performed using DFT calculations using the DMol3 software package as implemented by Materials Studio. DMol3 employs numerical radial functions centered at atomic positions as basis sets. The GGA base PBE functional was used to treat exchange correlational effects and the double numerical precision – DNP 3.5 basis set was used.  All energy calculations were deemed to be converged once an SCF convergence of \(10^{-6}\) was achieved. Additionally for the periodic systems a k-point mesh with \(0.03\;\mathring{A}^{-1}\) maximum spacing between mesh points to have the internal energies calculated to an accuracy of \(1\;m\,eV\).

The systems for which we require the internal energies to perform our calculations are a pure Li primitive unit cell, a LiH primitive unit cell and the hydrogen molecule. First the primitive unit cells of Li and LiH and also the hydrogen molecule were geometry optimized to obtain the optimal structures. The optimal structures of the Li unit cell (primitive unit cell of a BCC lattice with lattice constant = 3.0155 Å), LiH unit cell (primitive unit cell of a FCC lattice with lattice constant = 2.9409 Å) and the hydrogen molecule (H-H bond length = 0.748444 Å)  are shown in figures 1, 2 and 3 respectively. The internal energies of the systems of interest are shown in table 01.

Figure 1 : Primitive unit cell of Li. This is a primitive unit cell of a BCC lattice with lattice constant = 3.0155 Å.

Figure 2 : Primitive unit cell of LiH. This is a primitive unit cell of a FCC lattice with lattice constant = 2.9409 Å.

Figure 3 : Hydrogen molecule. Hydrogen bond length = 0.748444 Å.

Table 01 : Internal Energies of systems of interest.

SystemInternal Energy
\(\left(e\,V\right)\)
Li-204.660
LiH-221.233
H\(_2\)-31.678

Now using the primitive unit cells of Li and LiH and comparing their respective grand potentials, we see that \(\Omega_{Li}=\Omega_{LiH}\) or Li and LiH are thermodynamic equilibrium when,

\begin{equation}\mu_H^{eq}=E_{LiH}-E_{Li}=-16.573\;eV;\label{req_mu_H}\end{equation}

Or when, (using equation \eqref{muH2_H})

\begin{equation}\mu_{H_2}^{eq}=-33.146\;eV.\label{req_mu_H2}\end{equation}

Now in order to determine the phase diagram we need to determine the pressure at which \(\mu_{H_2}=-33.146\;eV\) for a given temperature. To do this we look at equation \eqref{H2_ideal_chem_pot}. From the NIST-JANAF Thermochemical Tables for molecular hydrogen we have obtained the plot , shown in figure 4, of \(\tilde{\mu}_{H_2}\left(T,p^0\right)\) vs \(T\) for a reference pressure \(p^0=0.1\;MPa\).

Figure 4 : Plot of \(\tilde{\mu}_{H_2}\left(T,p^0\right)\) vs \(T\) for a reference pressure \(p^0=0.1\;MPa\).

Now for a given temperature we can solve for the pressure \(p\) that makes \(\mu_{H_2}=\mu_{H_2}^{eq}\), by rewriting equation \eqref{H2_ideal_chem_pot} as,

\begin{equation}\ln\left(\frac{p^{eq}}{p^0}\right)=\frac{\mu_{H_2}^{eq}-E_{H_2}^{total}-\tilde{\mu}_{H_2}\left(T,p^0\right)}{k_BT}.\label{solve_for_p}\end{equation}

The curve \(\ln\left(\frac{p^{eq}}{p^0}\right)\) vs \(T\) which is the phase boundary between Li and LiH is shown in figure 5. For any point in phase space, corresponding to a point above and/or to the left of the \(\ln\left(\frac{p^{eq}}{p^0}\right)\) vs \(T\) curve, LiH is the thermodynamically preferred structure. Conversely, any point in phase space, corresponding to a point below and/or to the right of the \(\ln\left(\frac{p^{eq}}{p^0}\right)\) vs \(T\) curve, pure Li is the thermodynamically preferred structure.

Figure 5 : Plot of \(\ln\left(\frac{p^{eq}}{p^0}\right)\) vs \(T\).

 

 

 

Vibrational Modes of Water Calculated in the Harmonic Approximation

Introduction

In this post we will report on our efforts to determine the normal modes of atomic vibrations in a water molecule and compute the associated frequencies. For this we will explicitly calculate the mass-weighted Hessian matrix elements by calculating the change in energy when atoms are displaced by finite distances and thus numerically calculating the second derivatives of the energy of the water molecule with respect to the atomic coordinates (i.e. the Hessian matrix elements). We will compare our results for vibrational frequencies of normal modes in H\(_2\)O with the normal mode vibrational frequencies calculated by Materials Studio.

All energy calculations were performed using DFT calculations using the DMol3 software package as implemented by Materials Studio. DMol3 employs numerical radial functions centered at atomic positions as basis sets. The GGA base PBE functional was used to treat exchange correlational effects and the double numerical precision – DNP 3.5 basis set was used.  All energy calculations were deemed to be converged once an SCF convergence of \(10^{-6}\) was achieved.

In the Harmonic approximation, expanding the energy of the water molecule about the equilibrium positions is,

\begin{equation}E=E_0+\frac{1}{2}\sum_{i=1}^{9}\sum_{j=1}^{9}\left[\frac{\partial^2E}{\partial x_i\partial x_j}\right]_{\{x_i\}_{eq}}x_ix_j.\label{HarmonicApproximation}\end{equation}

Where, \(x_i\) is the \(\left(i-3\lfloor\frac{i-1}{3}\rfloor\right)\)-th coordinate of the \(\lceil\frac{i}{3}\rceil\)-th atom. Hence the sum over \(i\) runs to 9 for the 3 possible coordinates of the 3 atoms in H\(_2\)O.

The Hessian matrix elements are defined as,

\begin{equation}H_{ij}=\left[\frac{\partial^2E}{\partial x_i\partial x_j}\right]_{\{x_i\}_{eq}}.\label{Hessian}\end{equation}

The matrix form of the the equations of motion for the atomic vibrations can be written as, (according to  equation 5.7)

\begin{equation}\frac{d^2\boldsymbol{x}}{dt^2}=-\boldsymbol{Ax}.\label{EOM}\end{equation}

Where \(\boldsymbol{A}\) is the mass-weighted Hessian matrix whose matrix elements are given by,

\begin{equation}A_{ij}=\frac{1}{m_i}H_{ij}.\label{MassWeightedHessian}\end{equation}

Once A is determined the normal modes and corresponding vibrational frequencies can be determined using the eigenvectors {e\(_i\)} and eigenvalues {\(\lambda_i\)} respectively. The i-th eigenvector e\(_i\) define thes i-th normal mode of atomic vibrations and the corresponding eigenvalue (\(\lambda_i\)) defines the angular frequency of that vibrational mode by, \(\omega_i=\sqrt{\lambda_i}\).

Calculations

Figure 1: Relaxed H\(_2\)O molecule. The molecule has been oriented such that the O atom is at the origin of the coordinate system and the entire unperturbed molecule is in the xy-plane and symmetric about the yz-plane.

The notation \(\delta x_i=\delta r\) is used to denote an atomic displacement. Here \(\delta r\) gives the magnitude of the displacement, \(\alpha=\lceil \frac{i}{3} \rceil \) specifies the atom that is moving (\( \alpha \in \{1,2,3\}\) where the atoms are numbered as shown in figure 1) and \(\left(i-3\lfloor\frac{i-1}{3}\rfloor\right)\) denotes the Cartesian coordinate of the displacement. For example  \(\delta x_6=0.1\mathring{A}\) means the H atom assigned #2 has been moved \(0.1\;\mathring{A}\) in the z direction, from its equilibrium position.

The Hessian matrix elements were calculated using numerical differentiation for a finite displacements of the atomic positions.  In principle the numerical differentiation approaches the actual derivative for infinitesimally small displacements, however we need to consider the accuracy of our energy calculations and choose a displacement that is sufficiently large to produce a difference in energy between the unperturbed and perturbed molecule.

Table 1: Energy of a perturbed H\(_2\)O molecule and the corresponding change in energy from the unperturbed energy \(E_0\) for different \(\delta x_1\) values

\(\delta x_1\)
(Å)
\(E(\delta x_1) \)
(Ha)
\(E(\delta x_1)\) -\(E_0\)
(Ha)
0.02-76.3781430.000462
0.03-76.3775650.00104
0.04-76.3767540.001851
0.05-76.375710.002895
0.06-76.3744320.004173

Table 1 above lists the energy of a perturbed H\(_2\)O molecule for different \(\delta x_1\) values and the corresponding change in energy from the unperturbed energy \(E_0 = -76.378605 \)Ha. From the values in the table we decided to use 0.04 Å displacements to perturb the molecule in order to obtain a perturbation in energy that is \(>\;10^{-3}\;\mathring{A}\). For the rest of the post an atomic displacement will be specified by \(\pm\delta x_i\), since the magnitude of any displacement is always chosen to be 0.04 Å.

When evaluating the derivatives using numerical differentiation we chose to use the symmetric difference quotient (SDQ) method for improved accuracy. In this method the derivative of a function \(f(x)\) at \(x=x_0\) is calculated, (using a small displacement from \(x_0\) , h)

\begin{equation}f^{\prime}=\frac{f(x_0+h)-f(x_0-h)}{2h}.\label{SDQ}\end{equation}

The alternative methods to compute the derivatives are the forward difference quotient (FDQ) and the backward difference quotient (BDQ), which are defined in as,

\begin{equation}f^{\prime}=\pm\frac{f(x_0\pm h)-f(x_0)}{h}.\label{FDQBDQ}\end{equation}.

Where choosing the + ve sign gives the FDQ and – sign gives the BDQ.

Now to show that SDQ is the best choice; shown below in table 2 is the gradient of the total energy with respect to \(\delta \boldsymbol{x}_3\) evaluated at the equilibrium position using the SDQ, FDQ and BDQ. Since we are at equilibrium the actual gradient with respect to any coordinate should vanish by definition. None of our numerical results gradients vanish exactly but the gradients obtained using SDQ are consistently lower than the gradients from the other methods by atleast an order of magnitude, indicating this is the best choice to compute derivatives.

The gradient of the total energy with respect to \(\delta x_7\) evaluated at the equilibrium position using the SDQ, FDQ and BDQ.

\(\delta x_7\)
(Å)
\(E(+\delta x_7)\)
(Ha)
\(E(-\delta x_7)\)
(Ha)
\(\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\frac{dE}{dx_7}\)
\(\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\)(Ha/Å)
FDQBDQSDQ
0.02-76.378353-76.3783510.0126-0.0127-5E-05
0.03-76.37805-76.3780230.0185-0.0194-0.00045
0.04-76.377634-76.3775530.024275-0.0263-0.0010125
0.05-76.377111-76.3761640.02988-0.04882-0.00947
0.06-76.376485-76.3769350.035333333-0.0278333330.00375

With our choice to use the SDQ method to compute numerical derivatives, the Hessian matrix elements defined in equation \eqref{Hessian} become,

\begin{equation}H_{ij}=\left[\frac{\partial^2E}{\partial x_i\partial x_j}\right]_{\{x_i\}_{eq}}=\frac{E(\delta x_i +\delta x_j)+E(-\delta x_i -\delta x_j)-E(\delta x_i -\delta x_j)-E(-\delta x_i +\delta x_j)}{4 \delta x_i \delta x_j}.\label{Hessian_num}\end{equation}

Results

Table 3 below shows the Hessian matrix calculated using equation \eqref{Hessian_num}. The units of the matrix elements are \(Ha/\mathring{A}^2\). And the corresponding mass-weighted Hessian matrix is shown in table 4, with matrix elements having units of \(\frac{Ha}{amu.\mathring{A}^2}\).

Table 3. The Hessian of the water molecule, matrix elements have units of \(Ha/\mathring{A}^2\)

\(\delta x_1\)\(\delta x_2\)\(\delta x_3\)\(\delta x_4\)\(\delta x_5\)\(\delta x_6\)\(\delta x_7\)\(\delta x_8\)\(\delta x_9\)
\(\delta x_1\)2.3218752.22E-120-1.161093750.91968750.00E+00-1.16109375-0.91968750
\(\delta x_2\)2.22E-121.5582812500.7003125-0.779218750-0.7003125-0.779218750
\(\delta x_3\)000.0078125-2.22E-122.22E-12-0.0034375-2.22E-122.22E-12-0.0034375
\(\delta x_4\)-1.161093750.7003125-2.22E-121.26921875-0.80968750-0.107968750.112.22E-12
\(\delta x_5\)0.9196875-0.779218752.22E-12-0.80968750.744218752.22E-12-0.110.035468750
\(\delta x_6\)00-0.003437502.22E-120.00406252.22E-120-0.0009375
\(\delta x_7\)-1.16109375-0.7003125-2.22E-12-0.10796875-0.112.22E-121.269218750.80968750
\(\delta x_8\)-0.9196875-0.779218752.22E-120.110.0354687500.80968750.744218752.22E-12
\(\delta x_9\)00-0.00343752.22E-120-0.000937502.22E-120.0040625

Table 04 : The mass-weighted Hessian of the water molecule. [Units : \(\left(\frac{Ha}{amu.\mathring{A}^2}\right)\) ]

\(\delta x_1\)\(\delta x_2\)\(\delta x_3\)\(\delta x_4\)\(\delta x_5\)\(\delta x_6\)\(\delta x_7\)\(\delta x_8\)\(\delta x_9\)
\(\delta x_1\)0.145122631.39E-130-0.0725710810.0574826240.00E+00-0.072571081-0.0574826240
\(\delta x_2\)1.39E-130.0973962300.043771173-0.0487029980-0.043771173-0.0487029980
\(\delta x_3\)000.0004883-1.39E-131.39E-13-0.000214852-1.39E-131.39E-13-0.000214852
\(\delta x_4\)-1.1519930050.694823395-2.20E-121.259270513-0.8033411050-0.1071224820.1091378112.20E-12
\(\delta x_5\)0.912478917-0.7731111722.20E-12-0.8033411050.7383855052.20E-12-0.1091378110.0351907430
\(\delta x_6\)00-0.00341055702.20E-120.0040306582.20E-120-0.000930152
\(\delta x_7\)-1.151993005-0.694823395-2.20E-12-0.107122482-0.1091378112.20E-121.2592705130.8033411050
\(\delta x_8\)-0.912478917-0.7731111722.20E-120.1091378110.03519074300.8033411050.7383855052.20E-12
\(\delta x_9\)00-0.0034105572.20E-120-0.00093015202.20E-120.004030658

Now that we have the mass-weighted Hessian matrix, we calculate the eigenvalues and eigenvectors to determine the vibrational frequencies and corresponding normal modes. The eigenvalues are listed in table 5. The first three eigenvalues are non-negligible and are the vibrational frequencies of the normal modes of vibrations in a water molecule. The eigenvectors corresponding to these vibrational frequencies, which define the normal modes of these vibrational frequencies are listed in table 6. Inspecting the results table 6 we see that the normal modes given by \(\boldsymbol{e}_1\), \(\boldsymbol{e}_2\) and \(\boldsymbol{e}_3\) respectively depict the anti-symmetric stretching, symmetric stretching and bending normal modes shown respectively in the animated figures 2, 3 and 4 (click on figures for animated view).

Figure 2 : The Anti-symmetric stretching normal mode

Figure 3 : The Symmetric stretching normal mode

Figure 2 : The Bending normal mode

 

Table 05: Eigenvalues of the mass weighted Hessian matrix

i\(\lambda_i\) \((10^{29} \;s^{-2})\)
15.285202922
24.977711427
30.896378227
4-0.032981179
50.013024606
60.009386832
72.21E-07
80.000113396
93.56E-05

Table 06 : The normal modes corresponding to the non negligible eigenvalues of the mass-weighted Hessian matrix (i.e. the normal modes of vibration in H\(_2\)O)

\(x_i\)\(\boldsymbol{e}_1\)\(\boldsymbol{e}_2\)\(\boldsymbol{e}_3\)
x\(_1\)-0.069920136-1.19E-12-4.15E-13
x\(_2\)7.82E-13-0.049910964-0.070473846
x\(_3\)-7.67E-145.79E-144.52E-13
x\(_4\)0.555004088-0.5847203840.427667384
x\(_5\)-0.4353458970.3960511580.560907592
x\(_6\)1.31E-131.14E-128.48E-13
x\(_7\)5.55E-010.584720384-0.427667384
x\(_8\)0.4353458970.3960511580.560907592
x\(_9\)1.09E-12-2.20E-136.45E-12

Now we compute the vibrational frequencies of the normal modes and compare them with the vibrational frequencies calculated by Material Studio in table 7.

Table 07 : The vibrational frequencies (reported in inverse wavelength) of normal modes calculated in this work compared with the results from Materials Studio.

Normal ModeVibrational Frequency (cm\(^{-1}\))Discrepancy between results
\(\frac{\left|This\;work-Materials\;Studio\right|}{This\;work}\times100\)
(%)
This WorkMaterials Studio Results
Anti-symmetric stretching3859.403848.970.27
Symmetric stretching3745.443734.410.29
Bending1589.401615.691.65

As we can see from the results in table 7 our explicit calculation is in very good agreement with the results from Materials Studio. This is not surprising since Materials Studio employs a similar method to evaluate the Hessian matrix. In Materials Studio the Hessian matrix elements are obtained by displacing each atom in the system and computing a gradient vector, which is then used to build the second derivative matrix.

References

Surface Reconstruction of Pt(110) Surface

Introduction

The Pt(110) surface has been experimentally shown to undergo surface reconstruction. The mechanism for surface reconstruction is to have missing alternate rows of the top layer of the surface in a \((2\times1)\) surface unit cell. The aim of this study is to verify that the reconstructed Pt(110) is energetically favorable by computing and comparing the surface energy of Pt(110) surface with and without surface reconstruction

All energy calculations were carried out using plane-wave based DFT. The GGA based PBE functional was used to treat the exchange-correlation effects. The ion and core were treated using ultrasoft pseudopotentials generated on the fly (OTFG ultrasoft) with Koelling-Harmon relativistic treatments. Pseudo atomic calculations for Pt treated the 4f14 5s2 5p6 5d9 6s1 electrons as valence electrons and the electrons in lower levels were treated as part of the frozen core.

Calculation of Lattice constant for bulk Pt

Before attempting to compute surface energies we must verify we have the correct lattice parameters for the conventional unit cells used to build the desired surfaces. By correct lattice parameter(s) we refer to the lattice parameter(s) that minimizes the total energy calculated by our DFT code consistent with our choice of exchange correlation functional and pseudo potentials. Using any other lattice constant will introduce errors in our calculations by introducing artificial stress.

Figure 1 : (a) Conventional unit cell and (b) Primitive unit cell of bulk Pt.

Bulk Pt naturally occurs in the FCC crystal structure. The conventional unit cell and primitive unit cell of bulk Pt are shown in figure 1. The first step of our calculations was to geometry optimize the primitive unit cell of bulk Pt, allowing the lattice parameters to relax. For this calculation we used a cutoff of \(480\;eV\) and \(k\) point separation of \(0.04\;\mathring{A}^{-1}\). These calculation parameters were the default parameters for the ‘ultra-fine’ setting for CASTEP calculations as implemented by Materials Studio™. The geometry optimization yielded a lattice constant of \(3.967\;\mathring{A}\) for the conventional unit cell (\(2.805\;\mathring{A}\) for the primitive unit cell).

The next step in our investigation is to test the convergence of the total energy with respect to the cutoff energy and \(k\) point sampling. Figure 2 shows the plots for the results of the calculations to test for convergence. From these plots we see that total energy of the primitive unit cell converges to a tolerance of \(1\;m\,eV\) for cutoff energy of \(480\;eV\) and k point separation of \(0.025\;\mathring{A}^{-1}\).

Figure 2 : Plots of results used to test convergence of total energy with respect to (a) E\(_{cutoff}\)  and (b) k point sampling.

Calculation of surface energy of Pt(110) surface without surface reconstruction

The surface energy of Pt(110) was calculated by cleaving the optimized conventional unit cell of bulk Pt along the (110) plane. The cleaved unit cell was repeated infinitely in directions parallel to the cleaved plane and a finite number of times perpendicular to the cleaved plane to obtain a slab of desired thickness (in atomic layers). Both surfaces of the slab have Pt(110) surface. Now to employ plane wave based DFT calculations a vacuum slab unit cell was created containing a unit cell of the infinite slab (in directions parallel to the surface) and a vacuum layer in the direction normal to the surface. Figure 3 shows such a vacuum slab unit cell containing a 5 atomic layer thick slab separated by \(10\;\mathring{A}\) of vacuum from the next parallel slab.

Figure 3 : Two views of a vacuum slab unit cell containing a 5 atomic layer thick slab and a \(10\;\mathring{A}\) of vacuum layer.

Once we have defined a vacuum slab as depicted in figure 3, we fix the position of the bottom two layers to mimic bulk and allow the remaining layers of atoms to relax. In our DFT calculations for the slab unit cell, a self-consistent dipole correction was introduced to correct for the dipole rising from the asymmetry of the cell due to constraining the two bottom layers.

The surface energy (\(\sigma\)) can be calculated using,

\begin{equation} \sigma=\frac{ E_{slab}-n\,E_{bulk} }{A}. \label{surf_energy}\end{equation}

Where, \(E_{slab}\) is the total energy of the vacuum slab unit cell, \(E_{bulk}\) is the total energy of a primitive unit cell of Pt in bulk, \(A\) is the surface area of the slab in the vacuum slab unit cell and \(n\) is the number of Pt atoms in the vacuum slab unit cell.

First we need to determine the convergence tolerance of our calculation of  vacuum slab unit cell energy with respect to cutoff energy and k point sampling used. Here we chose to use the cutoff energy and k point separation used for the bulk unit cell, i.e. \(480\;eV\) and \(0.025\;\mathring{A}^{-1}\) respectively. Table 01 below lists the values obtained for the total energy of the vacuum slab unit cell (6 atomic layer slabs with vacuum separation of \(8\;\mathring{A}\)) for different cutoff and k point sampling around the chosen values. From this table we see that the total energy of the vacuum slab unit cell has a convergence tolerance of \(\sim\;10\;meV\).

Table 01 : Tests for convergence of total energy with respect to k point separation and cutoff energy

Cutoff Energy
(eV)
k point separation
(1/Å)
Total energy of vacuum slab
(eV)
4400.025-78303.015
4800.025-78303.017
5200.025-78303.017
4800.020-78303.009
4800.025-78303.017
4800.030-78303.003

For a given choice of cutoff energy and k point sampling, the thickness of the vacuum layer and the slab thickness will be the two parameters that effect the convergence of the surface energy. In order to test the convergence of the surface energy with respect to these parameters we compute \(\sigma\) while varying vacuum thickness (for a slab with 5 atomic layers) and also  \(\sigma\) while varying slab thickness (with a vacuum thickness of \(10\;\mathring{A}\)). The plots for these calculations are shown in figure 4. From the plots presented in figure 4 we can say for a vacuum thickness of \(8\;\mathring{A}\) and a slab thickness of 6 atomic layers the surface energy converges to (keeping in mind that the convergence tolerance of the total energy of the vacuum slab with respect to energy cutoff and k point sampling is \(10\;m\,eV\)),

\begin{equation}\sigma_{unreconstructed}=125\pm 1 \;\frac{m\,eV}{\mathring{A}^2}.\label{sigma_unrec}\end{equation}

Figure 4 : Surface energy of Pt(110) unreconstructed surface as function of (a) vacuum thickness and (b) slab thickness.

 

Calculation of surface energy of Pt(110) surface with surface reconstruction

Figure 5 : Vacuum slab unit containing the “missing row” reconstructed surface of Pt(110). Note the missing row of atoms in the top layer compared to the third layer.

The vacuum slab unit cell containing the reconstructed Pt(110) depicted in figure 5 was geometry optimized and the resulting total energy was used to calculate the surface energy of the reconstructed surface using \eqref{surf_energy}. Ideally we would want to retest the convergence tolerance with respect to cutoff energy, k point sampling, slab thickness and vacuum thickness. However given the computational cost of doing these calculations we chose to assume that the convergence tolerance for the surface energy is the same as surface energy calculation for the unreconstructed surface. With this assumption we get,

\begin{equation}\sigma_{reconstructed}=119\pm 1 \;\frac{m\,eV}{\mathring{A}^2}.\label{sigma_rec}\end{equation}

Conclusion

From the results given in equations\eqref{sigma_unrec} and \eqref{sigma_rec} we  see that,

\begin{equation}\sigma_{reconstructed}<\sigma_{unreconstructed}.\label{result}\end{equation}

This indicates that the “missing row” reconstructed surface of Pt(110) is energetically favorable compared to Pt(110) without surface reconstruction. This finding is in line with experimental evidence supporting this claim [1].

References

[1] H. Niehus, “Analysis of the Pt(110)–(1 × 2) surface reconstruction”, Surf. Sci., 145 (1984), pp. 407-418

Comparing ScAl in CsCl and NaCl Structures and Determining the Optimal Lattice Parameter of the Preferred Structure

The goal of this post is to identify if ScAl, which has AB stoichiometry, exists in the CsCl structure (figure 1) or NaCl Structure (figure 2 and figure 3). To determine this primitive cells of ScAl were produced for both types of structures. The plan is to plot the cohesive energy of the structure as a function of the volume per ScAl dimer. From these plots the optimal lattice parameters for each structure can be determined. Then by comparing the cohesive energies of the two structures with optimal lattice parameters we can determine which structure is preferred by ScAl. All energy calculations were carried out using plane-wave based DFT. The GGA based PBE functional was used to treat the exchange-correlation effects. The ion and core were treated using ultrasoft pseudopotentials generated on the fly (OTFG ultrasoft) with Koelling-Harmon relativistic treatments. Pseudo atomic calculations for Sc treated the 3s2 3p6 3d1 4s2 electrons as valence electrons and the electrons in lower levels were treated as part of the frozen core. Pseudo atomic calculations for Al treated the 3s2 3p1 electrons as the valence electrons.

Figure 1 : Primitive unit cell of ScAl in the CsCl structure. This a simple cubic structure with a two atom basis.

Figure 2 : (a) Conventional unit cell of ScAl in the NaCl structure. This a face centered cubic structure with a two atom basis. (b) Primitive unit cell of ScAl in the NaCl structure.

However before we begin calculating the data points to populate the plots we described in the previous paragraph, we need to select appropriate cutoff energy and \(\vec{k}\) point mesh for our calculations. The constraints on our choice of cutoff energy and \(\vec{k}\) point mesh are (1) computational cost, (2) convergence of results to a desired tolerance and (3) ensuring that we have approximately the same \(\vec{k}\) point density for each structure.

To determine the cutoff energy we use for our calculations we first determine the approximate lattice parameter of CsCl structure primitive unit cell. For this we use a cutoff energy of 460 eV and \(\vec{k}\) point mesh of \(8\times8\times8\) . The cutoff energy and \(\vec{k}\) point mesh chosen here are the default settings for the “ultra-fine” quality energy calculation using the CASTEP tool implemented in Material Studio. The cohesive energy vs volume per  ScAl dimer plot (figure 4) indicates that cohesive energy is minimized when the lattice parameter is \(\sim 3.5\;\mathring{A}\).

Figure 3 : Plot of Energy per ScAl dimer vs lattice parameter for ScAl in the CsCl structure, used to determine the approximate value for the optimal lattice parameter.

Next we investigated the convergence of the total energy of a primitive unit cell of ScAl in the CsCl structure with respect to the cutoff energy used for the calculation. The results were plotted as shown in figure 5.  From this plot note that we get a convergence of \(\sim\;1\;m\,eV\) for a cutoff energy of \(500\;m\,eV\).

Figure 4 : Calculated total energy of a unit cell of ScAl in the CsCl structure (lattice constant = 3.5 \(\mathring{A}\)) vs the cutoff energy used for the calculation.

Next to obtain the most suitable \(\vec{k}\) point mesh, we plot the energy per ScAl dimer for ae a fixed cutoff energy of \(500\;eV\) while varying the number of \(\vec{k}\) points used to sample the first brillouin zone. Since all the reciprocal lattice vectors ( and real space lattice vectors) have the same length, we can specify the \(\vec{k}\) mesh by specifying the number of \(k\)-points used along each reciprocal lattice vector. Figure 6 shows this plot; and we can see that a \(8\times8\times8\) \(\vec{k}\) point is sufficient for the energy per ScAl dimer to have a convergence of \(\sim\;10\;m\,eV\). The resulting spacing between sampled \(k\) points is \(0.0357\;\mathring{A}^{-1}\). To ensure our subsequent calculations have the same degree of convergence, we will impose that the separation between two adjacent \(k\) points that are sampled along a reciprocal lattice vector is at most \(0.0357\;\mathring{A}^{-1}\).

Figure 5 : Plot of energy per ScAl dimer vs volume per ScAl dimer.

Now we are ready to calculate the energy per ScAl dimer and the corresponding volume per dimer, for both structures and various lattice parameters. Figure 6 shows the plot of energy per ScAl dimer vs volume per dimer, for ScAl in the CsCl and NaCl structure. From the plots in figure 6 it is clear that ScAl prefers CsCl structure over the NaCl structure. From the best fit line we obtain the optimal lattice parameter in the CsCl structure to be \(3.38\;\mathring{A}\). If ScAl were to be found in the NaCl structure the optimal lattice parameter would be \(4.00\;\mathring{A}\).

Experimentally ScAl has been verified to exist in CsCl structure with a lattice parameter of \(3.450\;\mathring{A}\) [1]. Our results verify this and estimate the lattice parameter within \(\sim\;2%\) of the experimentally determined lattice constant.

 

[1] O. Schob and E. Parthe. Ab Compounds with Sc Y and Rare Earth Metals. I. Scandium and
Yttrium Compounds with Crb and Cscl Structure. Acta Crystallographica, 19:214-&, 1965.