JP 9943427

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

6998 J. Phys. Chem.

B 2000, 104, 6998-7011

Density Functional Theory Study of Proton Mobility in Zeolites: Proton Migration and
Hydrogen Exchange in ZSM-5

Jason A. Ryder,† Arup K. Chakraborty,*,†,‡,§ and Alexis T. Bell†,|


Chemical and Materials Sciences DiVisions, Lawrence Berkeley National Laboratory, and Departments of
Chemistry and Chemical Engineering, UniVersity of California, Berkeley, California 94720-1462
ReceiVed: December 9, 1999; In Final Form: April 18, 2000

Acidic protons in zeolites are known to be mobile at elevated temperatures. In this study, density functional
theory was used to identify the reaction pathways for proton migration in a model that represents the zeolite
ZSM-5. In the absence of water, the acidic proton “hops” or migrates between two of the four O atoms
surrounding an aluminum center with an activation barrier of 28 kcal/mol. During proton transfer, the O
atoms stretch closer together in order to stabilize the transition state. This is revealed by a 13.4° decrease in
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

the O-Al-O bond angle. Adsorbed water bridges the proton donor and acceptor sites, reducing the barrier
Downloaded via INDIAN INST OF TECH KANPUR on August 12, 2020 at 18:20:06 (UTC).

height by 24 kcal/mol. Hence proton migration depends heavily on the local geometry and conditions of the
zeolite. We show that experimentally undetectable amounts of water can greatly influence the measured rates
and apparent activation barriers. We broaden the scope of our study to consider hydrogen exchange with
other gas-phase species of the form RO-H (RO ) CH3O, CH3CH2O) and R-H (R ) H, CH3, C2H5, C3H7,
C6H5). It is evident that to a first approximation the activation barrier increases with an increase in the
polarizability of the species RO-H. For the chemical series R-H, the activation energy increases with the
deprotonation energy of the interacting species R-H. We also calculate the overall reaction rate constants
for proton hopping and hydrogen exchange.

I. Introduction interest. In solids, such as zeolites, this necessitates the use of


a truncated solid model or cluster approximation. Much work
Zeolites are microporous aluminosilicates formed from has been performed in the past decade to model the catalytically
oxygen-sharing silicate, [SiO4]4-, and aluminate, [AlO4]3-, active sites in zeolites using such a quantum mechanical cluster
tetrahedra. The presence of trivalent aluminum introduces a approach.4-28 Many of these studies employ molecular models
negative charge within the crystal structure. Counterions, such with imposed symmetry and relaxed geometry, approximations
as H+, adsorb to oxygen sites proximate to the aluminum designed for computational efficiency. Others restrict the cluster
centers, balancing the charge. The resulting acid sites are widely from relaxing in ways that are inconsistent with the real zeolite.
accepted as the catalytic site for hydrocarbon cracking, alky- Still others have carried out calculations with periodic boundary
lation, and isomerization reactions in zeolites. conditions and clusters embedded in a classical electrostatic field
Developing an understanding of the relationship between local representing the extended zeolite structure.
zeolite structure and catalytic activity remains an aim of Experimental studies suggest that an acidic proton is not fixed
heterogeneous catalysis and reaction engineering research. to a specific zeolitic oxygen site; rather, it “hops” or migrates
Current experimental methods provide an average picture of from one site to other neighboring sites.29-33 Variable temper-
the zeolite; hence, complete information on local structure and ature 1H magic angle spinning (MAS) NMR studies of
reaction mechanisms within the zeolite remains elusive. Com- dehydrated H-ZSM-5 zeolite provide evidence that acidic
putational methods have emerged as a powerful complementary protons migrate between the four oxygen atoms surrounding
tool to experimental methods.1-3 Using electronic structure the tetrahedral aluminum center in the following fashion
calculations, one can obtain information pertinent to the local
electronic and structural properties of the zeolite, and the
potential energy hypersurface characterizing interactions with (1)
host molecules. This assists the development of rational reaction
schemes and structure-function relationships. In addition, such
calculations allow the estimation of key quantities that are Since acidic protons are strongly bound, mobility is likely
experimentally observable (e.g., vibrational frequencies and restricted to this local intrasite hopping mechanism. Theoretical
overall reaction rate constants). calculations by Sauer et al.6,7 report an activation energy of 13
The computational expense of electronic structure calculations ( 3 kcal/mol for proton hopping in a 3T atom zeolite model,
is high, often restricting the size of a molecular model to only which is in reasonable agreement with experimental measure-
those atoms vital to capturing the chemical phenomena of ments of 4-11 kcal/mol.26-28 In this unconstrained symmetric
model zeolite, the O-Al-O angle closes from 94.2° at the
† Department of Chemical Engineering, University of California.
‡ reactant state to 74.7° at the transition state. The 20° change
Materials Science Division, Lawrence Berkeley National Laboratory.
§ Department of Chemistry, University of California. indicates that the flexibility of the O-Al-O angle plays an
| Chemical Science Division, Lawrence Berkeley National Laboratory. important role in the reaction coordinate of the proton-hopping
10.1021/jp9943427 CCC: $19.00 © 2000 American Chemical Society
Published on Web 07/06/2000
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 6999

Figure 1. Equilibrium structures (a, c) and transition state (b) for proton-hopping reaction.

reaction. This, in turn, makes us question whether the uncon- II. Theoretical Methods
strained geometry relaxation employed by Sauer et al.6,7 is
appropriate in this case. More recent contributions by Sauer et Quantum chemical calculations are performed on a model
al.8 on proton hopping in faujasite model zeolite report proton- zeolite system. We represent the protonated zeolite using a 34-
hopping barriers of 16.0, 23.5, 24.1, 25.3, 25.6, and 26.2 kcal/ atom cluster, as shown in Figure 1a. The cluster contains an Al
mol for jumps between different O-O pairs about the Al atom surrounded by shells of O and Si atoms. The terminal O
tetrahedron. In addition to employing a more extended repre- atoms are fixed in their crystallographic positions, as reported
sentation of the zeolite structure, embedded cluster techniques by Olson et al.;35 hence, no symmetry constraints were used
were used to take into account long-range crystal interactions. for any of the studied structures. Dangling bonds are terminated
Higazy et al.34 have suggested that adsorbed water assists by H atoms located 1.0 Å from each terminal O atom in the
proton transfer/mobility in H-ZSM-5. On the basis of measure- direction of the next T (tetrahedral) site, where a T site is an Al
ments of the ac conductivity of H-ZSM-5 at temperatures or Si atom. The overall geometry for the model is based on
below 373 K, they ascribe increases in the dielectric constant that of a T12 crystallographic site of ZSM-5. The anionic cluster
of the zeolite to water-assisted proton transfer. This provides is charge-compensated by a proton placed at one of the four O
an alternative mechanism for proton migration in zeolites, that atoms surrounding the Al atom.
of water-assisted proton migration: Geometry optimization calculations for minimum energy and
transition-state structures were performed using nonlocal, gradi-
ent-corrected density-functional theory (DFT).36 Calculations
were carried out to search for minima (i.e., reactants, products,
and adsorbed structures) and saddle points (i.e., transition-state
(2)
structures). Three different functionals were used to represent
the effects of exchange and correlation in the proton-hopping
reaction: Becke’s 3-parameter exchange with correlation func-
tionals of Lee, Yang, and Parr (B3LYP); Becke’s Half and Half
Equation 2 can be generalized to include any gas-phase exchange with correlation functionals of Lee, Yang, and Parr
hydrogen-bearing species in the zeolite channel that can function (BH&HLYP); and Becke’s 1988 exchange with correlation
as a proton shuttle in the zeolite. For example, simple alcohols functionals of Perdew and Wang (BPW91).37-41 The first two
of the form RO-H (e.g., H2O, CH3OH, CH3CH2OH, etc.) or methods employ hybrid functionals, which combine Hartree-
hydrogen and alkanes of the form R-H (e.g., H2, CH4, C2H6, Fock exchange with the Becke 1988 exchange functional.37 The
etc.) have been shown to participate in exchange with zeolites main difference between the two functionals lies in the weighting
at various temperatures. of exchange terms. B3LYP has become the functional of choice
In this work, density functional theory was used to investigate for calculating equilibrium structures for large systems.9 How-
the mobility of acidic protons in H-ZSM-5 zeolite. We consider ever, the BH&HLYP functional has been shown to give more
the migration of an acidic proton between two of four oxygen accurate determination of transition-state energies comparable
sites surrounding an aluminum center in a model zeolite. In to those attained from second-order Møller-Plesset perturbation
contrast to previous studies, we use a larger cluster model in theory (MP2).42 Since we are interested in transition states in
which the exterior atoms are constrained to preserve the this study, we rely primarily on the latter hybrid method.
geometric integrity of the zeolite. Using this constrained cluster Basis sets at the 6-31G or double-ζ level were used on all
formalism, we then examine the mechanism of local proton atoms. Diffuse and polarization functions were added to all
migration in H-ZSM-5 with and without the introduction of atoms, with the exception of terminal OH groups. While
gas-phase hydrogen-bearing species RO-H (RO ) OH, CH3O, computationally expensive due to the N3 scaling of the methods
CH3CH2O) and R-H (R ) H, CH3, C2H5, C3H7, C6H5). Using (where N is the number of basis functions), inclusion of diffuse
absolute rate theory, we provide an estimation the overall functions is vital to capturing the long-range behavior of
reaction rate constants for proton hopping and hydrogen hydrogen bonding within the zeolite. No corrections were made
exchange. Our results suggest that the experimental observations for basis-set superposition error. Zero-point energy corrections
for proton migration may be rationalized by the presence of were computed from the vibrational modes. All calculations
experimentally undetectable amounts of residual water in the were carried out using the Jaguar 3.5 suite of programs
zeolite. (Schrödinger, Inc.).43
7000 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

TABLE 1: Selected Bond Lengths (Å) and Angles (deg) of Equilibrium Structures and the Transition State for the
Proton-Hopping Reaction
(a) equilibrium structure (b) transition state (c) equilibrium structure

method DFT-BH&HLYP DFT-B3LYP HF DFT-BH&HLYP DFT-B3LYP HF DFT-BH&HLYP DFT-B3LYP HF


basis set 6-31G**++ 6-31G**++ DZP 6-31G**++ 6-31G**++ DZP 6-31G**++ 6-31G**++ DZP
cluster size 5T 5T 3T 5T 5T 3T 5T 5T 3T
reference this work this work 6, 7 this work this work 6, 7 this work this work 6, 7
OZ1-Al-OZ2 (deg) 96.9 96.2 94.2 83.6 83.3 74.7 96.1 95.6 94.2
r(OZ1-HZ) (Å) 0.97 0.97 0.9524 1.29 1.29 1.204 2.56 2.54
r(OZ2-HZ) (Å) 2.62 2.60 1.30 1.30 1.204 0.97 0.97 0.9524
r(Al-OZ1) (Å) 1.86 1.86 1.943 1.78 1.78 1.842 1.68 1.68 1.712
r(Al-OZ2) (Å) 1.68 1.68 1.712 1.77 1.78 1.842 1.86 1.85 1.943
EAct (kcal/mol) 32.3 25.8 13 ( 3
VTST (cm-1) 1785i 1520i

Vibrational modes of adsorbed complexes and transition-state four oxygen atoms surrounding the tetrahedral Al center. From
structures were computed for those atoms whose normal modes our zeolite cluster model, we choose two equilibrium oxygen
change most during reaction, and therefore contribute to the sites between which to probe the reaction pathway for proton
zero-point energy correction and prexponential factor. In Figure hopping. Although the protons can jump between all four
1 these include the proton donor and acceptor O atoms of the oxygen positions (for a total of six different jump paths), only
Brønsted acid site, denoted OZ1 and OZ2, respectively, the acidic one path has been studied so far. The sites are chosen because
proton, HZ, and any gas-phase or adsorbed species interacting of their accessibility to each other through the main channel of
with the site (e.g., R-H′). ZSM-5. The binding energy of a proton at these two sites differs
Overall reaction rate constants were computed using standard in energy by 3.2 and 3.6 kcal/mol for calculations using DFT-
statistical mechanics and absolute rate theory.44 We use the B3LYP and DFT-BH&HLYP, respectively.
harmonic approximation and include the contributions of the
The equilibrium structures and transition-state complex for
translational, rotational, and vibrational partition functions of
the proton-hopping reaction in H-ZSM-5 are shown in Figure
all gaseous species participating in the reaction and the
1. Parts a and c of Figure 1 correspond to local minima, whereas
vibrational contribution due to the zeolite cluster. Since the
Figure 1b is the structure of the transition state for proton
zeolite cluster is part of a solid, translational and rotational
partition functions for the zeolite were assumed to be equal in hopping between the two equilibrium states. We define the
the reactant and transition state. All molecular structures were activation energy for proton hopping, EAct, as the difference in
assumed to be in the ground-state electronic configuration. energy between the equilibrium structure and the transition state.
For the case of the proton-hopping reaction, the reaction rate These values, uncorrected for zero-point energy, are 32.3, 25.9,
constant can be expressed on a per site basis as and 20.5 kcal/mol for calculations using BH&HLYP, B3LYP,
and BPW91 density functionals, respectively. Zero-point energy

( )

kBT qTS,vib (ZPE) corrections reduce the reported energy barriers by 4.3
k) e-EAct/RT (3) kcal/mol. The imaginary frequency associated with the transition
h qH+OZ-,vib
state is 1785i cm-1, computed using BH&HLYP. A summary
where kB is Boltzmann’s constant, h is Planck’s constant, and of selected bond lengths and angles for local proton hopping in
EAct is the activation barrier including zero-point energy (ZPE) ZSM-5 is reported in Table 1.
corrections. In eq 3 we have written the vibrational partition
Figure 2 gives the energy along the reaction coordinate for
function for the transition-state structure and equilibrium
the proton-hopping reaction. Points along the minimum energy
Brønsted acid site as q‡TS,vib and qH+OZ-,vib, respectively. A path, defined as that corresponding to the steepest ascent along
similar expression can be written for the overall reaction rate
the potential energy hypersurface, were computed by optimizing
constant for hydrogen exchange between an adsorbed species,
the geometry of the cluster model with relaxation of the acidic
denoted RH′, and zeolite
proton restricted to the plane perpendicular to the line containing

( )

kBT qTS,vib atoms OZ1 and OZ2. Note that the greatest difference in energy
k) e-EAct/RT (4) computed using the three density functionals lies in the
h qRH′-H+OZ-,vib
transition-state region. This results from the different methods
Again the rate is calculated on a per site basis. Finally, we write of treating the exchange and correlation energy in each of the
the overall reaction rate constant for the hydrogen exchange functionals, which are not parametrized to yield accurate saddle
reaction between a hydrogen-bearing gas-phase species, such points (i.e., transition states). For reasons of comparison,
as dihydrogen or methane, and zeolite additional geometry optimization calculations were carried out

( )
‡ at the Hartree-Fock (HF) and MP2 levels of theory for the
NAVkBT qTS,vib
k) e-EAct/kBT (5) reactant, product, and transition state of the proton-hopping
h qRH′,transqRH′,rotqRH′,vibqH+OZ-,vib reaction. The activation energy computed using Hartree-Fock
where NA is Avogadro’s number and V is the molar volume of theory is the highest of all methods at 41.4 kcal/mol, consistent
the system. Note that we have included the translational, with the notion that HF theory overlooks the correlation energy.
rotational, and vibrational partition functions of the gas-phase A corresponding value of 35.7 kcal/mol is found using MP2
species. methods. The difference in energy barriers computed using MP2
and DFT-BH&HLYP methods is small, only 3.4 kcal/mol. The
III. Results and Discussion difference is more dramatic for the nonhybrid functional
Proton-Hopping Reaction. In the simplest case, we consider BPW91, which differs from the MP2 energy by 15.2 kcal/mol.
proton migration to proceed via short-range “hops” between the This agrees with previous studies that suggest hybrid functionals,
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7001

site to the acceptor site, thus artificially lowering the activation


barrier to proton hopping.
A comparison of our results with those reported by Sauer et
al.6,7 based on Hartree-Fock calculations is given in Table 1.
In contrast to the asymmetric, constrained cluster used in the
present study, Sauer et al. used a symmetric 3T atom model in
which all atoms were relaxed to optimize the transition-state
geometry. A notable difference in geometry of the two models
is found upon comparing the O-Al-O angle at the transition
state. Sauer et al. obtained a value of 74.7°, whereas our
calculations give 83.3-83.6°, a difference of 11°. Consistent
with this, Sauer et al. report an activation energy of 13 ( 3
kcal/mol for proton transfer. Though this value is closer to the
experimental results,29,30 we suggest that this is fortuitous since
complete relaxation of the O-Al-O angle seems inappropriate.
Our aim in restricting the external atoms of the model zeolite
cluster is to preserve the overall structure of the solid ZSM-5
while allowing flexibility to the interior region of the catalytic
site where the reaction occurs. We believe this is realistic
because reactions cause local geometric changes. We will
therefore explore other mechanisms for proton hopping shortly.
Figure 2. Energy versus reaction coordinate for proton-hopping To date, there are no direct experimental data on the flexibility
reaction. of the zeolite crystal framework. On the basis of the large
geometric and energetic differences seen from modeling proton
migration using a symmetric, relaxed cluster and asymmetric
constrained cluster, the effects of framework flexibility are
crucial for a full understanding of local zeolite chemistry.
Differences in effective pore dimensions computed from crystal-
lographic data and those deduced from molecular sieving
properties lend support to the thermally activated breathing
motion of the zeolite pores at elevated temperatures. These
breathing motions have been studied by Deem et al.45 for the
aperture O-O distances across the 10-membered ring in ZSM-
5. On the basis of a crystal dynamics approach, they found a
maximum standard deviation in the aperture O-O distances of
0.1901 Å at 550 K, where the mean pore diameter was taken
as 8.1011 Å. Given the relationship of pore diameter to the
circumference, πd, this 2.3% deviation corresponds to a 0.062
Å expansion/contraction between pairs of nearest-neighbor O
atoms due to the breathing motion. Our results show that the
O-O distance changes from 2.65 Å at the reactant state to 2.37
Å at the transition state, a difference of 0.28 Å. We conclude
that the concerted motion of the proton from the proton donor
to the proton acceptor O atom dominates proton hopping, and
Figure 3. O-Al-O angle versus reaction coordinate for proton- our simulation technique allows the appropriate degree of
hopping reaction. flexibility to the internal atoms to allow proton transfer.
Variable temperature 1H MAS NMR studies provide experi-
which incorporate exact exchange, perform better than nonhy- mental evidence of proton mobility in zeolites. Room temper-
brids for transition-state calculations. ature spectra give a spectrum with a narrow central line and
Figure 3 illustrates how the energy and the O-Al-O angle sharp spinning sidebands caused by interaction between the
vary as the reaction proceeds. During the proton-hopping acidic proton and the nearby aluminum nucleus.29-33 These
reaction, the O-Al-O closes by 13.4° from reactant state to spinning sidebands disappear at elevated temperatures along with
the transition state and then relaxes to reach the product state. a broadening of the central peak, indicating that acidic protons
This is uniform for all three density functionals and suggests become mobile. Experiments by Sarv et al.26 indicate proton
that the O atoms stretch closer together in order to stabilize the migration in dehydrated H-ZSM-5 (Si/Al ) 38) for tempera-
transition state. The constraints imposed on the exterior of the tures above 478 K. They determined the energy barrier for
cluster prevent complete relaxation of the O-Al-O angle. The proton hopping to be 11 kcal/mol. A similar study by Baba et
behavior of the O-Al-O angle suggests that proton hopping al.27 has estimated this barrier to be 4-5 kcal/mol for dehydrated
in zeolites is greatly facilitated by deformation of the zeolite H-ZSM-5 (Si/Al ) 21). It is worthwhile to note that a decrease
framework. Therefore, it would be inappropriate to carry out in the Si/Al ratio results in the increased mobility of the acidic
dynamic studies without explicit account of framework motion. proton, marked by a lower proton-hopping barrier. Both
Also, from the standpoint of kinetic studies, not imposing experimental values are considerably lower than that reported
constraints that prevent complete geometry relaxation (e.g., refs in this study; however, it is important to note that these
6 and 7) would allow easier access of the proton from the donor measurements represent an average over all protons in the
7002 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

Figure 4. Adsorption complexes (a, c) and transition state (b) for water-assisted proton-hopping reaction.

TABLE 2: Selected Bond Lengths (Å) and Angles (deg) of Adsorbed Complexes and the Transition State for Hydrogen
Exchange with Water
(a) adsorbed complex (b) transition state (c) adsorbed complex

method DFT-BH&HLYP MP2 DFT-B3LYP DFT-BH&HLYP MP2 DFT-BP86 DFT-BH&HLYP MP2 DFT-BP86
basis set 6-31G(**++) DZVP2 6-31G(**) 6-31G(**++) 6-31G(++) DZP/TZP 6-31G(**++) 6-31G(++) DZP/TZP
cluster size 5T 3T 3T 5T 3T 5T 5T 3T 5T
reference This work 10 11 this work 12 13 this work 12 13
OZ-Al-OZ2 (deg) 98.4 98.7 95.4 95.4 98.1 97.2 104.2
r(OZ1-HZ) (Å) 1.00 1.0267 1.039 1.47 1.36 1.32 1.93 1.81 1.71
r(OZ2-H′) (Å) 1.89 1.817 1.886 1.36 1.36 1.32 0.99 1.02 1.05
r(O′-HZ)(Å) 1.72 1.611 1.717 1.47 1.11 1.146 0.97 0.99 1.009
r(O′-H′) (Å) 0.98 1.0045 0.994 1.36 1.11 1.147 1.74 1.61 1.51
r(Al-OZ1) (Å) 1.83 1.934 1.879 1.76 1.69
r(Al-OZ2) (Å) 1.69 1.776 1.74 1.83
EAds (kcal/mol) -16.9 -13.5 -22.1 -16.0 -22.1 -22.0
EAct (kcal/mol) 6.9 6.2 2.0
VTST (cm-1) 762i 659i

zeolite. The present study focuses on the local structure of experimentally indistinguishable quantity of water dramatically
defect-free catalytic site under perfect vacuum. Several factors influences the observed kinetics of the proton-hopping reaction.
that could contribute to a discrepancy between experimental We will also discuss quantum tunneling effects in the last
values and those presented in this paper are noted here. section, when we present a summary of overall reaction rate
Site Defects. A change in the local geometry of the catalyti- constants and prexponential factors for proton hopping and
cally active site, as in the case of a site defect, would allow hydrogen exchange with H-ZSM-5.
more flexibility in the O-Al-O angle, lowering the activation Water-Assisted Proton Hopping. The adsorption of water
barrier for proton mobility about the defect site. Even a small to the Brønsted acid sites in zeolites is well characterized;
number of these sites could contribute to lowering the average however, its ability to mobilize acidic protons is not.46 Next
or apparent activation energy measured for proton hopping. we consider the role of water-assisted proton conduction in the
Assisted Proton Transfer. Small gaseous species such as water overall mechanism for proton migration. Figure 4 presents the
may act to participate in hydrogen exchange between the O adsorption complexes and transition-state structure for proton
atoms surrounding the aluminum center. At low concentration, exchange between water and the zeolite. Note that the 34-atom
this would be experimentally indistinguishable from the proton- cluster shown in Figure 1 has been used for all calculations
hopping reaction and lower the apparent activation barrier for presented in this paper. For descriptive purposes, we have
proton hopping. omitted the exterior portion of the cluster from the remaining
Quantum Tunneling. Protons participating in hydrogen- figures to focus attention on the local chemistry at the Brønsted
transfer reactions exhibit nonclassical character due to quantum acid site. The structures shown in Figure 4a,c correspond to
mechanical tunneling. This effect is most pronounced at low local minima on the PES. Each complex is composed of water
temperatures, where experimental rate constants are known to coordinated to the proton donor and acceptor sites by two
deviate from Arrhenius behavior. This deviation leads to lower hydrogen bonds. The adsorption energy of water to the zeolite,
measured values of the apparent activation barrier. defined as the difference in energy between gas-phase and
Site defects result from either the cleavage of the oxygen adsorbed water, EAds, is -16.9 kcal/mol for the structure shown
bridge between T sites or complete removal of a T site from in Figure 4a and -16.0 kcal/mol for the structure shown in
the zeolite framework proximate to the Brønsted acid site. For Figure 4c. As noted in Table 2, these values lie within the range
example, it has been proposed that dehydroxylation from reported in previous theoretical studies,11-16 -13.5 to -22.0
steaming results in the formation of defects in the form of silanol kcal/mol. The range of different values reported is a consequence
groups. Gonzales et al.10 found that introduction of a silanol of the quantum chemical method and zeolite model used in each
group immediately adjacent to a Brønsted acid site reduces the of the studies. The transition-state structure in Figure 4b consists
proton affinity by 13 kcal/mol. This loss of affinity combined of a hydronium ion coordinated to the proton donor and acceptor
with greater flexibility in the zeolite framework introduced by sites by two hydrogen bonds. This structure constitutes a first-
local defects could reduce the activation barrier for proton order saddle point whose negative force constant is assigned to
hopping. Because of the great diversity of possible defects, we the asymmetric stretch of water: the “rocking” mode of
will not consider this issue further in this paper. hydrogen exchange at the transition state. The frequency
We will consider the role of gas-phase species in the next associated with this rocking mode is 762i cm-1. The proton-
few sections. We will show that the presence of even an transfer energy, EAct, is 6.9 kcal/mol. ZPE corrections lower
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7003

TABLE 3: Calculated Overall Reaction Rate Constants for


the Dry and Water-Assisted Proton-Hopping Reactions
dry proton hopping water-assisted proton hopping
T (K) ko (s-1) k (s-1) ko (s-1) k (s-1) kH2O/kDRY
200 4.16E+12 1.12E-18 3.43E+12 2.05E+08 1.83E+26
298.15 6.15E+12 1.91E-08 4.08E+12 6.00E+09 3.13E+17
400 8.07E+12 4.18E-03 4.39E+12 3.39E+10 8.11E+12
500 9.80E+12 5.80E+00 4.53E+12 9.25E+10 1.60E+10
600 1.14E+13 7.35E+02 4.61E+12 1.80E+11 2.45E+08
700 1.29E+13 2.37E+04 4.67E+12 2.90E+11 1.22E+07
800 1.43E+13 3.25E+05 4.73E+12 4.16E+11 1.28E+06
900 1.56E+13 2.51E+06 4.81E+12 5.54E+11 2.21E+05
1000 1.68E+13 1.29E+07 4.88E+12 6.99E+11 5.39E+04

this barrier by 3 kcal/mol. Selected bond lengths and angles


for hydrogen exchange between water and zeolite are listed in
Table 2.
Previous theoretical studies support the presence of a hydro-
nium ion as a transition state for the hydrogen exchange reaction.
Zygmunt et al.14 have modeled hydrogen exchange of water
with ZSM-5 using a symmetric, relaxed 3T atom zeolite model.
Using MP2 methods, they computed the proton-transfer energy
to be 6.2 kcal/mol. This is in good agreement with our results Figure 5. Overall rate of proton hopping versus temperature for dry
prior to ZPE correction. Krossner et al.15,16 have reported an and water-assisted proton-hopping reaction.
activation barrier of 2.0 kcal/mol based on DFT calculations
carried out with the BP86 functional for an asymmetric, relaxed NMR experiment, we can compare the rates of proton hopping
5T cluster. Again, a major difference between the present and for different amounts of water over a range of temperatures.
previous studies is the allowed flexibility of the cluster. Our Figure 5 shows the overall rate of proton hopping versus inverse
calculations predict a value of 98.7° for the O-Al-O angle at temperature for a system containing 0.1 g of zeolite (Si/Al )
the transition state, compared to 96.8 and 95.4° for Zygmunt et 16) and 1 cm3 of vapor volume. Each line corresponds to a
al. and Krossner et al., respectively. Table 2 compares the results different molar ratio of water to zeolite, φ ) NH2O/NZH. In the
of our study with those reported previously. high molar ratio limit, φ ) 1, the system behaves as though all
The overall reaction rate constants and frequency factors for sites are saturated by water at low temperature, and hence the
dry and water-assisted proton hopping in the model zeolite were rate of proton mobility is equal to that of water-assisted proton
computed over a range of temperatures and are included in Table transfer. In the low molar ratio limit, φ ) 0, the system behaves
3. The large difference in rate between the two mechanisms, as though no sites contain adsorbed water at any temperature
on the order of 1026 at low temperature and 104 at high and the rate of proton migration is that characteristic of proton
temperature, emphasizes the large impact of even a small hopping in the absence of H2O. The more interesting physical
amount of water on the mobility of acidic protons about the cases lie between these extremes, where a vanishingly small
Brønsted acid site. To illustrate this further, we consider the amount of water is present in the system over a range of
overall rate of proton hopping, rHOP, to consist of a linear temperatures.
combination of rates of dry and water-assisted proton hopping,
Note that for temperatures e500 K, the activation barrier for
denoted rDRY and rH2O, respectively
proton migration, taken as the slope of ln(rHOP) versus 1/T,
rHOP ) (1 - θ)rDRY + θrH2O (6) resembles that of water-assisted proton transfer for all nonzero
values of φ. This corresponds to the physical case where some
of the water molecules present in the system are adsorbed on
where θ is the fractional coverage of water over the total number
Brønsted acid sites and assist the proton-hopping reaction.
of catalytically active sites defined by the Langmuir isotherm
Because of the large difference in magnitude of the two rates,
KH2OPH2O water-assisted proton hopping dominates the rate. Even as little
θ) (7) as 1 ppm of water per acidic proton, an amount indistinguishable
1 + KH2OPH2O by most experimental means, will reduce the apparent activation
energy from 28 kcal/mol for completely dry H-ZSM-5 to ∼4
In eq 11, PH2O is the partial pressure of water in atmospheres kcal/mol.
and KH2O is the equilibrium constant for water adsorbed to the Over the intermediate range of temperatures, 600-800 K,
Brønsted acid site computed using the formula water begins to desorb from the acid sites and the rate of proton
qH2O-H+OZ-,vib hopping diminishes. Beyond 800 K, almost all water in the
K H2 O ) e-EAds/RT (8) system resides in the gas phase, hence the rate of proton hopping
qH2O,transqH2O,rotqH2O,vibqH+OZ-,vib equals that of dry proton hopping. The temperature dependence
of the rate in the region above 600 K is nonlinear due to a shift
We have assumed an equilibrium distribution of water because in the water-zeolite equilibrium. The nonlinear behavior, if
the time scale associated with the diffusion of water in a unit measured judiciously and plotted over this temperature regime,
cell is 2 orders of magnitude smaller than the characteristic time would be interpreted as a negative activation barrier for proton
for proton hopping. hopping. However, if one assumed that the zeolite sample was
If we fix the total moles of zeolite in a system of constant completely dry and considered the first-order rate dependence
volume, as in the case of a sealed ampule used in a typical over the entire range of temperatures (e.g., refs 29 and 30), this
7004 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

Figure 6. Adsorption complexes (a, c) and transition state (b) for hydrogen exchange with methanol.

TABLE 4: Selected Bond Lengths (Å) and Angles (degrees) of Adsorbed Complexes and the Transition State for Hydrogen
Exchange with Methanol
(a) adsorbed complex (b) transition state (c) adsorbed complex

method DFT-BH&HLYP DFT-BLYP DFT-B3LYP DFT-BH&HLYP DFT-BP86 MP2 DFT-BH&HLYP DFT-BP86 MP2
basis set 6-31G(**++) DZVP2 6-31G(**) 6-31G(**++) DZPV DZP/TZP 6-31G(**++) DZPV DZP/TZP
cluster size 5T 3T 3T 5T 3T 5T 5T 3T 5T
reference this work 10 11 this work 16 18 this work 16 18
OZ1-Al-OZ2 (deg) 98.1 90.5 99.1 94.98 96.8 98.1 97.19 98.0
r(OZ1-HZ) (Å) 1.00 1.035 1.049 1.26 1.348 1.366 1.92 1.786 1.762
r(OZ2-H′) (Å) 1.86 1.837 1.51 1.348 1.366 1.00 1.052 1.049
r(O′-HZ) (Å) 1.66 1.569 1.482 1.01 1.123 1.086 0.97 1.005 0.993
r(O′-H′) (Å) 0.97 1.001 0.967 1.12 1.123 1.086 1.65 1.527 1.453
r(C-O′) (Å) 1.41 1.46 1.432 1.42 1.464 1.444 1.42 1.449 1.431
r(Al-OZ1) (Å) 1.83 1.945 1.88 1.73 1.864 1.808 1.70 1.788 1.728
r(Al-OZ2) (Å) 1.69 1.779 1.76 1.863 1.808 1.82 1.926 1.905
EAds (kcal/mol) -17.1 -16.7 -21.0 -15.9 -17.5 -20.9
EAct (kcal/mol) 6.5 2.6 0
VTST (cm-1) 397i 3561i

would lead to the result of a measured activation energy between previous studies.12,13,17-20 The transition-state structure in Figure
that of dry proton hopping and that of water-assisted proton 6b is composed of a methoxonium ion coordinated to the zeolite
transfer. framework through the proton donor and proton acceptor O
We have shown that the presence of an adsorbed water atoms. It is a first-order saddle point on the PES with one
molecule at the Brønsted acid site can dramatically influence negative force constant assigned to the asymmetric stretch of
the kinetics of proton hopping in H-ZSM-5. We suggest that methanol and hydrogen exchange. The imaginary frequency
this is an important reason for the observed low activation associated with this rocking mode is 397i cm-1. The computed
barriers for proton hopping in ZSM-5. We now broaden the activation energy is 6.5 kcal/mol. ZPE corrections reduce the
scope of our study to investigate the influence of other activating activation barrier by 2.6 kcal/mol. Selected bond lengths and
species on the mobility, and hence reactivity, of acidic protons. angles for hydrogen exchange between methanol and zeolite
Hydrogen Exchange with R-OH. In the last section we are given in Table 4.
examined the impact of adsorbed water on the kinetics of proton Other theoretical studies of methanol in zeolites have
mobility about the Al tetrahedron. On the basis of the similar concluded that the methoxonium ion is a transition state for
hydroxyl moiety in methanol and ethanol, we now consider the hydrogen exchange. DFT-BP86 calculations by Blaszkowski et
proton conduction properties of the general assisted proton- al.18 computed the activation energy to be 2.6 kcal/mol for a
hopping reaction (9), where RO-H represents the series of symmetric, fully relaxed 3T zeolite model. Bates et al.19 reported
simple alcohols H2O, CH3OH, and CH3CH2OH.
1.6 kcal/mol for proton transfer in a similar cluster using MP2
methods. Haase et al.20 determined that the barrier height was
negligible for MP2 calculations on an asymmetric, relaxed 5T
cluster model. The lower activation energies in both of these
(9) studies can be ascribed to the use of unconstrained clusters,
which allow full relaxation of the zeolite framework to stabilize
the transition state. Our calculations predict a value of 99.1°
for the O-Al-O angle at the transition state, compared to 95.0,
The adsorption complexes and transition-state structure for 96.0, and 96.8° for Blaszkowski et al., Bates et al., and Haase
hydrogen exchange between methanol and the zeolite are shown et al., respectively. A detailed comparison of the present work
in Figure 6. As in the case of water-assisted proton hopping, to the previous studies is included in Table 4.
each structure consists of methanol coordinated to the proton Figure 7 illustrates the different adsorption complexes for
donor and acceptor sites through the protons involved in transfer. proton exchange between ethanol and the zeolite. Parts a and c
The adsorption energy of methanol to the zeolite, EAds, is -17.1 of Figure 7 show ethanol coordinated to the acidic proton by
kcal/mol for the structure shown in Figure 6a and -15.9 kcal/ two hydrogen bonds. The calculated adsorption energies for the
mol for the structure shown in Figure 6c, consistent with two structures are -14.9 and -16.4 kcal/mol, respectively. The
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7005

Figure 7. Adsorption complexes (a, c) and transition state (b) for hydrogen exchange with ethanol.

TABLE 5: Selected Bond Lengths (Å) and Angles (deg) of Adsorbed Complexes and the Transition State for Hydrogen
Exchange with Ethanol
(a) adsorbed complex (b) transition state (c) adsorbed complex
method DFT-BH&HLYP DFT-BH&HLYP DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**++)
cluster size ST 5T 5T
OZl-Al-OZ2 (deg) 99.8 99.6 98.1
r(OZ1-HZ) (Å) 0.98 1.42 1.93
r(OZ2-H′) (Å) 1.88 1.58 1.00
r(O′-HZ) (Å) 1.75 1.01 0.97
r(O′-H′) (Å) 0.96 1.06 1.62
r(C-O′) (Å) 1.40 1.44 1.42
r(Al-OZ1) (Å) 1.83 1.73 1.70
r(Al-OZ2) (Å) 1.68 1.75 1.82
EAds (kcal/mol) -14.9 -16.4
EAct (kcal/mol) 7.5
transition-state structure in Figure 7b consists of ethoxonium Figure 8 portrays the hydrogen exchange reaction between
ion coordinated to the zeolite framework by two hydrogen dihydrogen and the zeolite cluster. Parts a and c of Figure 8
bonds. The activation energy of this complex is 7.5 kcal/mol. show free, dissociated molecular hydrogen in proximity of the
Selected bond lengths and angles for hydrogen exchange Brønsted acid site. No energy minima were found for dihydro-
between ethanol and zeolite are provided in Table 5. gen adsorbed to either acid site of the zeolite cluster. Figure 8b
Hard-soft acid-base theory (HSAB) predicts that soft acids shows the activated complex of dihydrogen with the acidic
will prefer soft bases and hard acids will prefer hard bases. A proton. This corresponds to a saddle point in the PES with one
measure of softness of an acid or base is its polarizability. From negative force constant associated with proton transfer. The
a physical standpoint, water, methanol, and ethanol are polariz- imaginary frequency attributed to the transition state is 1679i
able molecules. This is illustrated by the index of refraction cm-1. The activation energy, defined as the difference in energy
exhibited by these compounds: 1.333, 1.3288, and 1.3611, between the activated complex and reactants, EAct, is 31.9 kcal/
respectively. It is evident that to a first approximation the mol. The ZPE correction raises the activation barrier by 1.5
activation energy for hydrogen exchange with R-OH increases kcal/mol. Selected bond lengths and angles for the hydrogen
with an increase in polarizability, denoted by a larger index of exchange reaction with dihydrogen are listed Table 6 along with
refraction of the species RO-H. We propose that since the a comparison with previous related efforts.
proton is a hard acid, it prefers a hard base. Hence the more Deuterium exchange reactions on H-ZSM-5 have been
polarizable the base becomes (as indicated by a higher refractive performed by Biscardi et al.47 They found that D2 dissociates
index), in the order CH3OH < H2O < CH3CH2OH, the higher and exchanges with surface hydrogen atoms between 600 and
the activation barrier for assisted proton transfer about the 1000 K, and then H desorbs from the surface as HD and H2.
Brønsted acid site. The prexponential factor and activation energy were computed
Hydrogen Exchange with R-H. The utility of zeolites in by fitting the HD desorption peak. These values are 4.16 × 108
cracking and other hydrocarbon re-forming processes motivates cm3/(Al site‚s) and 21 kcal/mol, respectively. Our computed
discussion of the mobility of acidic protons due to hydrogen barrier, corrected for ZPE, is 12 kcal/mol higher than the
exchange with participating reactant and product species. This experimental value. In view of this result, we make reference
involves the series dihydrogen, methane, and higher order to arguments made previously in the paper. The experimental
hydrocarbons, all of which are rich in hydrogen and have been value provides an average view of hydrogen exchange with all
shown to participate in exchange with zeolites at various available hydrogen on the zeolite surface. This includes defect
temperatures. The reaction proceeds in the following fashion: sites with lower proton affinity, hence more ready exchange.
Figure 9 shows proton exchange reaction between methane
and H-ZSM-5. Parts and and c of Figure 9 depict free methane
in proximity of the Brønsted acid site. No stable structure was
(10)
found for the adsorption complex of methane with the Brønsted
acid site. Figure 9b shows the activated complex of methane
In eq 10 R-H denotes the reacting species (R ) H, CH3, C2H5, with the zeolite cluster. This consists of a carbonium species,
C3H7, C6H5). CH5+, associated with the proton donor and acceptor sites
7006 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

Figure 8. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with dihydrogen.

Figure 9. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with methane.

TABLE 6: Selected Bond Lengths (Å) and Angles (deg) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with Dihydrogen
(b) activated complex
(a) dissociated species (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP MP2 DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**++) 6-31G(**++)
cluster size 5T 5T IT 5T
reference this work this work 20 this work
OZ-Al-OZ2 (deg) 96.9 97.9 90.1 96.1
r(OZ1-HZ) (Å) 0.97 1.16 1.09
r(OZ2-H′) (Å) 1.17 1.09 0.97
r(H-HZ) (Å) 1.05 1.22 0.74
r(H-H′) (Å) 0.74 1.05 1.22
r(Al-OZ1) (Å) 1.86 1.77 1.88 1.68
r(Al-OZ2) (Å) 1.68 1.76 1.88 1.86
EAct (kcal/mol) 31.9 32.0
VTST (cm-1) 1679i

through the two protons involved in the exchange. The transi- one negative force constant assigned to the rocking mode of
tion-state structure for hydrogen exchange corresponds to a proton exchange. The frequency associated with this mode is
saddle point on the potential energy surface with one negative 1147i cm-1. The computed activation energy for hydrogen
force constant assigned to the rocking mode of carbonium and exchange is 40.0 kcal/mol. ZPE corrections reduce the observed
proton exchange. The imaginary frequency of this mode is 1435i barrier by 1.7 kcal/mol. Selected bond lengths and angles for
cm-1. The calculated activation barrier, uncorrected for zero- hydrogen exchange between ethane and zeolite and a compari-
point energy, is 40.7 kcal/mol. ZPE corrections reduce the son with previous related efforts are provided in Table 8.
observed barrier by 2.3 kcal/mol. Experimental studies by Unlike methane and ethane, propane provides the opportunity
Larson et al.48 determined the activation energy for deuterium to probe two chemically different carbon environments within
exchange to be 33.4 kcal/mol for exchange of CD4 with the same hydrocarbon molecule. These include the methyl
H-ZSM-5. This result lies within 4 kcal/mol of our ZPE carbon (C′) and methylene carbon (C′′), as shown below
corrected barrier. Selected bond lengths and angles for minimum
energy and transition structures are provided in Table 7, along
with a comparison with previous related studies. (11)
Parts a and c of Figure 10 depict free ethane in proximity of
the Brønsted acid site. No stable structure was found for the
adsorption complex of ethane with the Brønsted acid site. Figure
10b shows the activated complex of ethane with the zeolite
cluster. This consists of one carbon of ethane associated with Figure 11 shows the proton exchange reaction between a
the proton donor and acceptor sites through the two protons methyl group of propane and the zeolite. Parts a and c of Figure
involved in the exchange. The transition-state structure corre- 11 depict free propane in proximity of the Brønsted acid site.
sponds to a saddle point in the potential energy surface with No stable structure was found for the adsorption complex of
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7007

Figure 10. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with ethane.

Figure 11. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with methyl group of propane.

TABLE 7: Selected Bond Lengths (Å) and Angles (degrees) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with Methane
(b) activated complex
(a) dissociated species (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP DFT-BP86 DFT-BH&HLYP DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**) 6-31G(**++) 6-31G(**++)
cluster size 5T 5T 3T 3T 5T
reference this work this work 22 23 this work
OZ1-A1-OZ2 (deg) 96.9 95.7 92.3 96.1
r(OZ1-HZ) (Å) 0.97 1.41 1.399 1.322
r(OZ2-H′) (Å) 1.41 1.399 1.322 0.97
r(C-HZ) (Å) 1.28 1.282 1.322 1.09
r(C-H′) (Å) 1.09 1.28 1.282 1.322
r(Al-OZ1) (Å) 1.86 1.75 1.801 1.68
r(Al-OZ2) (Å) 1.68 1.75 1.801 1.86
EAct (kcal/mol) 40.7 31.3 37.7
VTST (cm-1) 1435i 1128i 1832i

TABLE 8: Selected Bond Lengths (Å) and Angles (deg) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with Ethane
(b) activated complex
(a) dissociated species (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP DFT-BP86 DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**) 6-31G(**++)
cluster size 5T 5T 3T 5T
reference this work this work 21 this work
OZl-Al-OZ2 (deg) 96.9 95.6 88.8 96.1
r(OZ1-HZ) (Å) 0.97 1.49 1.330
r(OZ2-H′) (Å) 1.47 1.330 0.97
r(C′-HZ) (Å) 1.26 1.323 1.09
r(C′-H′(Å) 1.09 1.28 1.319
r(Al-OZl) (Å) 1.86 1.75 1.854 1.68
r(Al-OZ2) (Å) 1.68 1.75 1.853 1.86
EAct (kcal/mol) 40.0 28.2
VTST (cm-1) 1147i 1374i
propane with the Brønsted acid site. Figure 11b shows the point in the potential energy surface with one negative force
activated complex of propane with the zeolite cluster. This constant assigned to the rocking mode of proton exchange. The
consists of the methyl carbon associated with the proton donor frequency associated with this mode is 1142i cm-1. The
and acceptor sites through the two protons involved in the computed activation energy is 40.5 kcal/mol. Zero-point energy
exchange. The transition-state structure corresponds to a saddle corrections lower the barrier by 1.3 kcal/mol. Selected bond
7008 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

Figure 12. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with methylene group of propane.

TABLE 9: Selected Bond Lengths (Å) and Angles (deg) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with the Methyl Group of Propane
(a) dissociated species (b) activated complex (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**++)
cluster size 5T 5T 5T
OZ1-Al-OZ2 (deg) 96.9 95.6 96.1
r(OZ1-HZ) (Å) 0.97 1.50
r(OZ2-H′) (Å) 1.46 0.97
r(C′-HZ) (Å) 1.24 1.09
r(C′-H′) (Å) 1.09 1.29
r(Al-OZl) (Å) 1.86 1.75 1.68
r(Al-OZ2) (Å) 1.68 1.75 1.86
EAct (kcal/mol) 40.5
VTST (cm-1) 142i

TABLE 10: Selected Bond Lengths (Å) and Angles (deg) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with the Methylene Group of Propane
(a) dissociated species (b) activated complex (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**++)
cluster size 5T 5T 5T
OZ-A1-OZ2 (deg) 96.9 96.1 96.1
r(OZ-HZ) (Å) 0.97 1.55
r(OZ2-H′) (Å) 1.47 0.97
r(C′′-HZ) (Å) 1.24 1.09
r(C′′-H′) (Å) 1.09 1.30
r(A1-OZ1) (Å) 1.86 1.76 1.68
r(A1-OZ2) (Å) 1.68 1.74 1.86
EAct (kcal/mol) 39.2
VTST (cm-1) 1029i
lengths and angles for hydrogen exchange between propane and methylene group. Although in spirit this is the opposite of our
zeolite are provided in Table 9. predicted trend, the experimental deviation reported more than
Figure 12 shows the proton exchange reaction between the accounts for the differences in the two measured values.
methylene group of propane and zeolite. Parts a and c of Figure As a final case in a series of hydrogen exchange reactions,
12 depict free propane in proximity of the Brønsted acid site. we consider the activation of benzene by the acidic proton in
Figure 12b shows the activated complex of propane with the zeolite. Figure 13 shows the hydrogen exchange reaction
zeolite cluster. This consists of the methylene carbon associated between benzene and zeolite. Parts a and c of Figure 13 depict
with the proton donor and acceptor sites through the two protons free benzene in proximity of the Brønsted acid site. No stable
involved in the exchange. The transition-state structure corre- structure was found for the adsorption complex of benzene with
sponds to a saddle point in the potential energy surface with the Brønsted acid site. Figure 13b shows the activated complex
one negative force constant assigned to the rocking mode of of benzene with the zeolite cluster. This consists of benzene
proton exchange. The frequency associated with this mode is associated with the proton donor and acceptor sites through the
1029i cm-1. The computed activation energy is 39.2 kcal/mol. two protons involved in the exchange. The transition-state
Zero-point energy corrections lower the barrier by 0.7 kcal/ structure corresponds to a saddle point in the potential energy
mol. Selected bond lengths and angles for hydrogen exchange surface with one negative force constant assigned to the rocking
between the secondary carbon of propane and zeolite are mode of proton exchange. The imaginary frequency of this mode
provided in Table 10. is 840i cm-1. The computed energy barrier is 33.6 kcal/mol.
In situ 1H MAS NMR experiments by Stepanov et al.49 ZPE corrections reduce the observed activation energy by 6.6
measured the activation energy for hydrogen exchange between kcal/mol. Selected bond lengths and angles for hydrogen
deuterated propane and the acidic protons in H-ZSM-5. For exchange between benzene and zeolite are provided in Table
the methyl group, they measured a value of 26 ( 2 kcal/mol, 11 along with a comparison with previous related studies.
whereas they found a higher value of 28 ( 2 kcal/mol for the Experimental measurements of the activation of benzene in
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7009

Figure 13. Dissociated species (a, c) and activated complex (b) for hydrogen exchange with benzene.

TABLE 11: Selected Bond Lengths (Å) and Angles (deg) of Dissociated Species and the Activated Complex for Hydrogen
Exchange with Benzene
(b) activated complex
(a) dissociated species (c) dissociated species
method DFT-BH&HLYP DFT-BH&HLYP DFT-BLYP DFT-BH&HLYP
basis set 6-31G(**++) 6-31G(**++) 6-31G(**) 6-31G(**++)
cluster size 5T 5T 3T 5T
reference this work this work 25 this work
OZl-Al-OZ2 (deg) 96.9 97.5 96.1
r(OZl-HZ) (Å) 0.97 1.49 1.48
r(OZ2-H′) (Å) 1.44 1.49 0.97
r(C′-HZ) (Å) 1.24 1.23 1.09
r(C′-H′) (Å) 1.09 1.21 1.22
r(Al-OZ) (Å) 1.86 1.75 1.81 1.68
r(Al-OZ2) (Å) 1.68 1.74 1.80 1.86
EAct (kcal/mol) 33.6 21.2
VTST (cm-1) 840i

H-ZSM-5 yield a value of 14.4 kcal/mol, significantly below


the 27.0 kcal/mol ZPE-corrected barrier predicted for our
constrained cluster model zeolite.28 Again, the experimental
value provides an average view of hydrogen exchange with all
available hydrogen on the zeolite surface, whereas our model
considers only one reactive site.
On the basis of the results of our electronic structure
calculations, we seek to develop some qualitative relation
between the structure of the molecular species R-H and its
ability to exchange hydrogen with the Brønsted acid site. Note
that in each of Figures 8-12, the mechanism for hydrogen
exchange between R-H and the Brønsted acid site requires
lengthening of the R-H′ bond in order to reach the transition
state of the reaction. In the case of dihydrogen, the H-H′ bond
length stretches from 0.74 Å at the reactant state to 1.05 Å at
the transition state. For methane, the CH3-H′ bond lengthens
from 1.09 Å at the reactant state to 1.28 Å at the transition
state. Similar trends are seen for the remaining hydrocarbons
in the series R-H.
We propose that some relationship lies between the weaken-
ing of the R-H′ bond and the ability of a species R-H to effect Figure 14. Deprotonation energy versus activation energy for hydrogen
exchange with R-H.
hydrogen exchange. We define the energy due to the deproto-
nation of a proton from a species R-H, EDP, as its ability to the MP2/6-31G**++ level of theory. It is evident that to a first
donate a proton approximation the activation barrier, EAct, increases with an
increase in the deprotonation energy, EDP, of the activating
R-H f R- + H+ EDP (12) species. One interpretation of this trend is that the early part of
the reaction coordinate is dominated by weakening of the R-H′
Figure 14 gives a plot of the deprotonation energy versus bond. One measure of the ability of this bond to be weakened
activation energy for the series of species R-H (R ) H, CH3, is the deprotonation energy of the species. The greater the ability
C2H5, C3H7, C6H5) considered in this study. The values of EDP of a species to donate protons, the lower the activation barrier
were computed for the appropriate ionic and neutral species at for hydrogen exchange.
7010 J. Phys. Chem. B, Vol. 104, No. 30, 2000 Ryder et al.

model experimental system, we have shown that an experimen-


tally undetectable quantity of water can significantly affect the
kinetics of proton migration. Thus, water-assisted proton transfer
provides a kinetically favorable alternative mechanism for proton
migration about the Brønsted acid site.
The kinetics of hydrogen exchange in zeolite varies with the
identity of the hydrogen-bearing chemical species. Through the
systematic application of density functional theory and the
constrained cluster method to a series of hydrogen exchange
reactions, we are able to obtain a qualitative relationship between
the activation energy for exchange and chemical properties of
the gas-phase species. Also, a comparison of computed exchange
rates to that of dry proton hopping give a measure of the
feasibility of exchange as an alternative pathway for protons to
migrate about the Al tetrahedron.

Acknowledgment. This work was supported by the Office


of Industrial Technology of the U.S. DOE under contract number
DE-AC03-SF763098. Additional support was provided by a
graduate research fellowship from the National Science Founda-
tion. Computational resources were provided by the National
Figure 15. Computed reaction rate constants versus temperature for Energy Resource Supercomputer Center (NERSC). We thank
hydrogen exchange with R-H. Mark Rice, Dr. François Gilardoni, and Dr. John Nicholas for
very thoughtful and stimulating conversations.
Rate Constants for Hydrogen Exchange. Figure 15 shows
the dependence of the reaction rate constants, k, on temperatures References and Notes
for the each of the hydrogen exchange reactions involving R-H. (1) Neurock, M.; van Santen, R. A. Catal. Today 1999, 50, 445-450.
As expected, those species with the lowest activation barriers (2) Bates, S. P.; van Santen, R. A. AdV. Catal. 1998, 42, 1-114.
(e.g., H2, C6H6) boast the highest rates for hydrogen exchange. (3) Nicholas, J. B. Top. Catal. 1997, 38, 377-390.
Note that for those species whose activation barriers are similar (4) Van Santen, R. A. Catal. Today 1997, 38, 157-171.
(5) Redondo, A.; Hay, P. J. J. Phys. Chem. 1993, 97, 11754-11761.
in magnitude (the series CH4, C2H6, C3H8) the prexponential (6) Sauer, J.; Kölmel, C. M.; Hill, J. R.; Ahlrichs, R. Chem. Phys. Lett.
factor determines the difference in the computed rate. 1989, 164, 193-198.
Additional calculations were performed to estimate the (7) Sauer, J. In Modeling of Structure and ReactiVity in Zeolites;
Catlow, C. R. A., Ed.; Academic Press: San Diego, CA, 1992.
contribution of quantum mechanical tunneling to the overall (8) Sauer, J.; Sierka, M.; Haase, F. In Transition State Modeling for
reaction rate constant for hydrogen hopping and hydrogen Catalysis; ACS Symposium Series 721; Truhlar, D. G.; Morokuma, K.,
exchange. Tunneling corrections were computed by fitting the Eds.; American Chemical Society: Washington, DC, 1999; p 359.
(9) Zygmunt, S. A.; Mueller, R. M.; Curtiss, L. A.; Iton, L. E. J. Mol.
energy minima, transition-state energy, and transition-state Struct. 1998, 430, 9-16.
frequency with an asymmetric one-dimensional Eckart func- (10) Gonzales, N. O.; Bell, A. T.; Chakraborty, A. K. J. Phys. Chem. B
tion.50 On the basis of our calculations, the largest contribution 1997, 101, 10058-10064.
of proton tunneling to proton hopping and hydrogen exchange (11) Rice, M. J.; Chakraborty, A. K.; Bell, A. T. J. Phys. Chem. B 1998,
102, 7498-7504.
is seen at temperatures below 400 K. At conditions of interest, (12) Gale, J. D. Top. Catal. 1996, 3, 169-194.
tunneling does not appear to be an important factor. For a more (13) Greatbanks, S. P.; Hillier, I. H.; Burton, N. A. J. Chem. Phys. 1996,
rigorous treatment of quantum tunneling effects, we recommend 105, 3770-3776.
the semiclassical tunneling approximations that require energy, (14) Zygmunt, S. A.; Curtiss, L. A.; Iton, L. E.; Erhardt, M. K. J. Phys.
Chem. 1996, 100, 6663-6671.
gradient, and Hessian information along the minimum energy (15) Krossner, M.; Sauer, J. J. Phys. Chem. 1996, 100, 6199-6211.
path.51 These methods demand a much more detailed under- (16) Jobic, H.; Tuel, A.; Krossner, M.; Sauer, J. J. Phys. Chem. 1996,
standing of the potential energy hypersurface but provide a more 100, 19545-19550.
(17) Gale, J. D.; Catlow, C. R. A.; Carruthers, J. R. Chem. Phys. Lett.
accurate estimate of the contribution of proton tunneling to the 1993, 216, 155-161.
overall reaction rate. (18) Blaszkowski, S. R.; van Santen, R. A. J. Phys. Chem. 1995, 99,
11728-11738.
(19) Bates, S.; Dwyer, J. J. Mol. Struct. 1994, 112, 57-65.
IV. Conclusions (20) Haase, F.; Sauer, J. J. Am. Chem. Soc. 1995, 117, 3780-3789.
(21) Allavena, M.; Kassab, E.; Evleth, E. J. Mol. Struct. 1994, 325, 85-
Using density functional theory, we have considered several 93.
different pathways for proton migration about the Brønsted acid (22) Evleth, E. M.; Kassab, E.; Sierra, L. R. J. Phys. Chem. 1994, 98,
site in model H-ZSM-5. In the absence of water, the acidic 1421-1426.
(23) Blaszkowski, S. R.; Nascimento, M. A. C.; van Santen, R. A. J.
proton “hops” or migrates between two of the four O atoms Phys. Chem. 1996, 100, 3463-3472.
surrounding an aluminum center with an activation barrier of (24) Blaszkowski. S. R.; Jansen, A. P. J.; Nascimento, M. A. C.; van
28 kcal/mol. During proton transfer, the O atoms stretch closer Santen, R. A. J. Phys. Chem. 1994, 98, 12938-12944.
together in order to stabilize the transition state. This is revealed (25) Kramer, G. J.; van Santen, R. A. J. Am. Chem. Soc. 1995, 117,
1766-1776.
by a 13.4-degree decrease in the O-Al-O bond angle. This (26) Truong, T. N. J. Phys. Chem. B 1997, 101, 2750-2752.
points to the importance of framework flexibility. The presence (27) Esteves, P. M.; Nascimento, A. C.; Mota, J. A. J. Phys. Chem. B
of adsorbed water bridges the proton donor and acceptor sites 1999, 103, 10417-10420.
during proton transfer, reducing the apparent activation barrier (28) Beck, L. W.; Xu, T.; Nicholas, J. B.; Haw, J. F. J. Am. Chem. Soc.
1995, 117, 11594-11595.
by 24 kcal/mol and increasing apparent rates by factors of 1026 (29) Sarv, P.; Tuherm, T.; Lippmaa, E.; Keskinen, K.; Root, A. J. Phys.
at low temperature and 104 at high temperature. Employing a Chem. 1995, 99, 13763-13768.
Proton Mobility in Zeolites J. Phys. Chem. B, Vol. 104, No. 30, 2000 7011

(30) Baba, T.; Komatsu, N.; Ono, Y. J. Phys. Chem. B 1998, 102, 804- (41) Becke, A. D. Phys. ReV. A 1988, 38, 3098-3100.
808. (42) Truong, T. N.; Duncan, W. T.; Bell, R. L. In Chemical Applications
(31) Ernst, H.; Freude, D.; Mildner, T.; Pfeifer, H. In Proceedings of of Density-Functional Theory; Laird, B. B., Ross, R. B., Ziegler, T., Eds.;
the 12th International Zeolite Conference; Materials Research Society: American Chemical Society: Washington, DC, 1996.
Warrendale, PA, 1999; p 2955. (43) Jaguar 3.5, Schrödinger, Inc., Portland, OR, 1998.
(32) Hunger, M. Catal. ReV. 1997, 39, 345-393.
(33) Bonn, M.; Bakker, H. J.; Domen, K.; Hirose, C.; Kleyn, A. W.; (44) McQuarrie, D. A. Statistical Mechanics; HarperCollins Publisher
van Santen, R. A. Catal. ReV. 1998, 40, 127-173. Inc.: New York, 1973.
(34) Higazy, A. A.; Kassem, M. E.; Sayed, M. B. J. Phys. Chem. Solids (45) Deem, M. W.; Newsam, J. M.; Creighton, J. A. J. Am. Chem. Soc.
1992, 53, 549-554. 1992, 114, 7198-7207.
(35) Olson, D. H.; Kokotallo, G. T.; Lawton, S. L.; Meier, W. M. J. (46) Ison, A.; Gorte, R. J. J. Catal. 1984, 89, 150-158.
Phys. Chem. 1981, 85, 2238-2243. (47) Biscardi, J. A.; Meitzner, G. D.; Iglesia, E. J. Catal. 1998, 179,
(36) Parr, R. G.; Yang, W. Density-Functional Theory of Atoms and 192-202.
Molecules; Oxford University Press: Oxford, U.K., 1989. (48) Larson, J. G.; Hall, K. H. J. Phys. Chem. 1965, 69, 3080-3089.
(37) Becke, A. D. J. Chem. Phys. 1993, 98, 1372-1377.
(38) Becke, A. D. J. Chem. Phys. 1993, 98, 5648-5652. (49) Stepanov, A. G.; Ernst, H.; Freude, D. Catal. Lett. 1998, 54, 1-4.
(39) Lee, C.; Yang, W.; Parr, R. G. Phys. ReV. B 1988, 37, 785. (50) Eckart, C. Phys. ReV. 1930, 35, 1303.
(40) Perdew, J. P. In Electronic Structure of Solids; Ziesche, P., Eschrig, (51) Duncan, W. T.; Bell, R. L.; Truong, T. N. J. Comput. Chem. 1998,
H., Eds.; Adademie Verlag: Berlin, 1991. 19, 1039-1052.

You might also like