Lec Notes MEGHalle
Lec Notes MEGHalle
Lec Notes MEGHalle
net/publication/251191601
CITATIONS READS
6 915
5 authors, including:
41 PUBLICATIONS 364 CITATIONS
Pacific Northwest National Laboratory
158 PUBLICATIONS 5,924 CITATIONS
SEE PROFILE
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
Natural gas conversion to aromatics over zeolite supported molybdenum catalysts View project
All content following this page was uploaded by Maria-Elena Grillo on 06 September 2015.
Preface
VI Preface
Contents
1 Introduction
............................................................... 1
VIII Contents
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
List of Contributors
Dr. M. Lüders
M. Schillinger
Daresbury Laboratory,
Institut für Theorie der Konden-
Daresbury, Warrington,
sierten Materie,
Cheshire.
Universität Karlsruhe
WA4 4AD. UK
D-76128 Karlsruhe
m.lueders@dl.ac.uk
matthias@tkm.physik.uni-karlsruhe.de
Prof. Dr. K. Busch
Department of Physics and School Dr. A. Klaedtke
of Optics: CREOL & FPCE, Advanced Technology Institute
University of Central Florida, School of Electronics and Physical
Orlando, FL 32816, Sciences
U.S.A. University of Surrey
kbusch@physics.ucf.edu GU2 7XH
United Kingdom
F. Hagmann a.klaedtke@surrey.ac.uk
Institut für Theorie der Konden-
sierten Materie, Dr. J. Hamm
Universität Karlsruhe Advanced Technology Institute
D-76128 Karlsruhe School of Electronics and Physical
Sciences
hagmann@tkm.physik.uni-karlsruhe.de
X List of Contributors
List of Contributors XI
1 Introduction
1 Introduction 3
Part II
10 M. E. Grillo et al.
2.1 Introduction
A catalyst is defined as a substrate capable to change the kinetics of a chem-
ical reaction increasing its rate by lowering the energy barrier to activate
the reactant molecules [1]. Catalysts are not consumed during a chemical
reaction, and their activity is defined as the reaction rate for conversion of
reactants into products. Understanding at a molecular level the relation be-
tween the catalyst surface composition and its activity, assists in the catalyst
design in industrial chemical processes.
Fig. 2.1. Schematic representation of energy profile for a catalyzed and uncat-
alyzed reaction.
typically done using Cartesian coordinates. Only recently, the first use of de-
localized internal coordinates for periodic systems was reported [2, 3]. This
is because use of internal coordinates for solids is more challenging than for
molecules. First, one needs to define a set of necessary primitive internal coor-
dinates such as bonds, angles, torsions that span the entire space of the solid.
For an infinite crystal, the number of internal coordinates is, in principle,
infinite; however, the number of degrees of freedom required for optimization
is just 3N − 3, where N is the number of the atoms in a single unit cell.
Therefore, only a subset of all internal coordinates of the crystal can be cho-
sen for geometry optimization.
where the index α runs over a translationally unique set of primitive internal
coordinates, i runs over all 3N Cartesian coordinates of the atoms in one unit
cell and the sum over R runs over all unit cells in the crystal. B is M × 3N
dimensional giving the rate of change of primitive internal coordinate qα on
periodic displacement of the i-th atom in the unit cell. The infinite sum over
atoms in Equation 2 does not pose any problems in practice, because each
primitive internal depends at most on the position of four atoms in the crys-
tal. Hence, the terms in sum over R vanish after only a few cells.
12 M. E. Grillo et al.
We can therefore obtain the matrix U that is needed to construct the de-
localized coordinates by diagonalizing the matrix F, which is much smaller
than G, since for solid-state systems M can be much greater than 3N . This
procedure will yield 3N − 3 vectors in U, which define our non-redundant
delocalized coordinates. The non-redundant delocalized internal coordinates
cnp of the system can be written as a linear combination of primitive internals
through X
cnp = qα Uαp (2.4)
α
Fig. 2.2. Crystal of graphite. The internal coordinates of this crystal do not span
the optimization space
are 3-5 times more efficient than Cartesian coordinates. The efficiency of
internal coordinates increases with system size, and they are particularly
beneficial in the study of systems that can undergo significant rearrangement
of geometry. Periodic systems that belong to that category include surface
reactions, molecular crystals and zeolites. For example, our recent study on
Al-substituted chabazite revealed that optimization in delocalized internals
converges in 21 cycles, whereas Cartesian coordinates require 96 cycles [2].
DFT periodic slab calculations show that the di-σ adsorption configu-
ration (with the ring bound through the olefinic C=C group) is preferred
over the atop (η 1 ) adsorption mode. The calculated adsorption energy of 76
kJ/mol is in good agreement to the measured energy of 90 kJ/mol determined
from ultra-high vacuum temperature programmed desorption (UHV-TPD)
[9]. The di-σ adsorption configuration was found starting from the slightly
distorted η 1 mode. Optimization in delocalized internals took 34 cycles to
obtain the di-σ structure, whereas optimization using Cartesian coordinates
14 M. E. Grillo et al.
Fig. 2.3. η 1 (a) and di-σ (b) adsorption modes of maleic anhydride on Pd(111).
Fig. 2.4. Reactant di-σ (a), transition state (b) and the stable intermediate (c) for
the C-O bond breaking in maleic anhydride on Pd(111) surface.
Figure 2.4 shows the reactant di-σ (a), transition state (b), and stable in-
termediate (c) structures for this reaction. The barrier of 59 kJ/mol is much
smaller than for the ring opening of the (η 1 ) structure. The ring opening of
di-σ structure is also an endothermic reaction, and leads to a stable interme-
The transition state (TS) search algorithm [10] implemented in our DFT
codes CASTEP and DMol3 is a generalized scheme based on the traditional
linear and quadratic synchronous transit (LST/QST) method [10] coupled
with previously-proposed conjugate gradient refinement ideas [11]. We refer
the interested reader to Ref. 10 and references therein for a complete descrip-
tion of the method.
The method can be summarized with a flow diagram (Figure 2.5). The
energies of suitable reactant and product structures are first calculated. This
is followed by a search for the LST maximum using essentially the method
of Halgren and Lipscpomb [10]. This maximum energy structure provides an
upper bound for the transition state. A conjugate gradient (CG) refinement
is initiated for further refinement of the estimated transition state, searching
in directions conjugate to the vector connecting reactants to products. CG
methods make intelligent use of gradient information, thereby requiring no
explicit Hessian to be calculated. As in a conventional optimization, the cal-
culation is considered converged if all the residual forces on the structure fall
below a certain threshold.
Fig. 2.5. Flow diagram illustrating the transition state search scheme.
16 M. E. Grillo et al.
If this is not achieved before the number of conjugate directions has been
explored, a QST maximum search is performed using the reactant, product,
and latest CG geometry. A new CG refinement cycle is started following the
QST maximum search. This is continued until the required convergence is
attained. A frequency analysis on this converged structure should yield ex-
actly one negative eigenmode, corresponding to the direction by which the
system would evolve away from the saddle point.
A number of calculations [12, 13, 14, 15] have been performed using the
new transition state search algorithm. For instance, it was used to investi-
gate reaction pathways in the methanol to gasoline conversion process over
zeolite catalysts [12, 13]. Reaction paths and energy barriers for the carbon-
oxygen (C-O) bond cleavage and first carbon-carbon (C-C) bond formation
were explored using all-electron periodic supercell calculations with DMol3 .
Calculations along these lines can help understand mechanism of a reaction
and ultimately improve existing catalysts.
The cleavage of the C-O bond of ethanol was studied by calculating the
reaction path for the methylation of a zeolitic oxygen at the aluminosilicate
Brønsted acid site. A reaction barrier of ∼ 54 kcal/mol was calculated for
such reaction. This is a concerted SN2-type reaction involving breaking of the
C-O bond in methanol and bond formation between the C atom and surface
oxygen, with the formation of a water molecule resulting from the transfer
of a zeolite proton to the hydroxyl group. A lower activation barrier of 44
kcal/mol was found in the presence of a second methanol molecule. The pres-
ence of a second methanol molecule in the zeolite cage causes the spontaneous
deprotonation of the zeolite, leading to the formation of the methoxonium
ion [12, 13].
18 M. E. Grillo et al.
and return a trajectory that contains at least one point in the vicinity of
the new minima, in this case M1 or M2. Points are shown in Figure 2.6 as
blue (minima) and red (maxima). In general, the path will also contain a few
points that do not correspond to stationary points (green). A conventional
NEB is intended to start from the end points and yield the TS and the entire
reaction pathway, i.e., many green points. In contrast, the algorithm imple-
mented in DMol3 is intended to start at the TS, and to locate the alternative
stationary points in the direction of reactants and products.
Fig. 2.7. Calculated transition state structure for the Diels-Alder reaction of bu-
tadiene and ethylene to cyclohexene.
Fig. 2.8. Resulting graph of the NEB calculation, and geometry of this new min-
imum structure. A new minimum is found at path coordinate 0.23. Note that the
cis-butadiene exhibits a bent structure. The green curve is the energy vs QST path
1. The yellow, blue, pink, and magenta curves denote the energy vs MEP paths.
The graph in Fig. 2.8 shows that the NEB procedure has found a new
minimum on the reaction path, at path coordinate 0.23, where the path co-
ordinate is defined between the reactants (0.0) and the products (1.0). The
new minimum structure shown in Fig. 2.8 exhibits a reactant complex in
which the cis-butadiene is bent rather than planar as in the reactant input
structure. This new minimum corrects the activation energy to 23.9 kcal/mol
20 M. E. Grillo et al.
Table 2.1. Calculated unit-cell edges (a0 , c0 ), volume per oxide formula-unit (V0 ),
internal fractional coordinates (z(Rh), x(O)). Percentage errors are reported in the
second line for each quantity.
CASTEP was also used to calculate the full elastic constant tensor. The
calculated elastic constants and bulk modulus obtained as a linear combi-
nation of elastic constants are summarized in Table 2.2. CASTEP uses the
finite strain technique to predict elastic constants, which applies a given ho-
mogeneous deformation (strain) and calculates the resulting stress. The bulk
modulus obtained by this technique of 2.29 Mbar agrees well with that ob-
tained by fitting the energy data to the Birch-Murnaghan EOS of 2.11 Mbar.
The band structure and density of states reveal that corundum rhodium
is a Mott-Hubbard insulator (see Fig. 2.9) with a metal-to-metal charge tran-
sition dominated by the dt2g (valence band edge) and deg metal (conduction)
at an energy of 1.4 eV. This is consistent with the observed semiconducting
behaviour of Rh2 O3 [24]. The band structure presents three groups of bands:
the low lying bands at about 20 eV correspond to oxygen 2p states, the group
of bands between 8 and 4 eV originates mainly from oxygen 2p states, the
bands from about 3 eV up to the Fermi level are of mainly Rh-4dt2g charac-
ter, and the conduction band consists of mainly oxygen 2p states hybridized
with Rh-4deg levels.
22 M. E. Grillo et al.
Fig. 2.9. Band structure and density of states (DOS). Red, green and blue curves
represent the p-, d- and s-DOS, respectively. The energies are relative to the Fermi
level in eV.
2.6 Summary
Industrial research in new materials development or improvement of existing
ones for catalysis applications, crystal growth (e.g. chemical vapour deposi-
tion), and microelectronics are commonly based upon phenomenological the-
oretical models, and high throughput experimentation. Recent advances in
quantum technology provide the materials researcher with tools to thought-
fully develop intelligent descriptors based upon the molecular understanding
of the underlying processes. A novel optimization technique for solid-state
calculations has been presented, which shows a superior performance for com-
plex problems, e.g. surface reactions. A powerful algorithm to optimize the
reaction path of a process by DFT within the transition state theory has
been described. This method allows not only the evaluation of the energy
difference between the particle at a minimum and the saddle point of the po-
tential energy surface (energy barrier), but also, it validates a transition state
by connecting it to the actual reactant and product. The constant pressure
optimization algorithm in CASTEP allows calculating the equation of state,
and the finite strain technique to predict the elastic properties for any kind
of existing or novel materials.
Acknowledgment
The authors are grateful to Professor Mathew Neurock for his advice in the
project on dissociation of maleic anhydride on Pd(111).
References
1. B.C. Gates. In: Catalytic Chemistry, (John Wiley & Sons 1992).
2. J. Andzelm, R.D. King-Smith, G. Fitzgerald: Chem. Phys. Lett. 335, 321
(2001).
3. K. Kudin, G.E. Scuseria, H.B. Schlegel: J.Chem.Phys. 114, 2919 (2001).
4. J. Baker, A. Kessi, B. Delley: J. Chem. Phys. 105, 192 (1996) .
5. J. Andzelm, G.Fitzgerald, N.Govind, D. King-Smith: J. Chem. Phys., submit-
ted (2003).
6. B. Delley,: J. Chem. Phys. 92, 508 (1990); J. Phys. Chem. 100, 6107 (1996) .
7. E. B. Wilson, J.C. Decius and P.C. Cross. In: Molecular Vibrations, (McGraw-
Hill, New York 1955).
8. V. Pallassana, M. Neurock, G. Coulston: J. Phys. Chem. 103, 8973 (1999).
9. Xu, Goodman, Langmuir 12, 1807 (1996).
10. T.A. Halgren, W.N. Lipscomb: Chem. Phys. Lett. 49, 225 (1977).
11. S. Bell, J.S. Crighton: J. Chem. Phys. 80, 2464 (1984); S. Fischer, M. Karplus:
Chem. Phys. Lett. 194, 252 (1992); J.E. Sinclair, R. Fletcher: J. Phys. C 7,
864 (1974).
12. N. Govind, J.W. Andzelm, K. Reindel, G. Fitzgerald: Int. J. Mol. Sci. 3, 423
(2002).
13. J. Andzelm, N. Govind, G. Fitzgerald, A. Maiti: Int. J. Quant. Chem. 91, 467
(2003).
14. A. Maiti, N. Govind, P. Kung, D. King-Smith, J. Miller: C. Zhang, G. Whitwell,
J. Chem. Phys. 117, 8080 (2002).
15. M. Petersen: Comp. Mater. Sci (submitted for publication).
16. G. Henkelman, G. Johannesson, H. Jonsson: Progress in Theoretical Chemistry
and Physics, ed. by S.D. Schwartz, (Kluwer Academic Publishers 2000).
17. K.N. Houk, R.J. Loncharich, J.F. Blake, W.L Jorgensen: J. Am. Chem. Soc.
111, 9172 (1989).
18. G.L. Kellog: J. Catal. 92, 167 (1985); G.L. Kellog: Surf. Sci. 171, 359 (1986).
19. J.P. Perdew, K. Burke, M. Ernzerhof: Phys. Rev. Lett. 77, 3865 (1996).
20. M.D. Segall, P.L.D. L indan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J.
Clark, M.C. Payne: J. Phys.: Cond. Matt. 14, 2717 (2002).
24 M. E. Grillo et al.
21. F. Birch: J. Geophys. Res. 57, 227 (1952); F.D. Murnaghan: Proc. Natl. Acad.
Sci. U.S.A. 30, 224 (1944).
22. C.E. Boman: Acta Chem. Scannd. 24, 116 (1970); K.R. Poeppelmeier, J.M.
Newsam, J.M. Brown: J. Solid State Chem. 60, 68 (1985).
23. L.W. Finger and R.M. Hazen, J. Appl. Phys. 51, 5362 (1980).
24. G. Bayer and H.G. Wiedemann, Thermochimica Acta 15, 213 (1976).