Vous êtes sur la page 1sur 7

DFT Study of Methanol Conversion to

Hydrocarbons in a Zeolite Catalyst


JAN ANDZELM, NIRANJAN GOVIND, GEORGE FITZGERALD,
AMITESH MAITI
Accelrys Inc., 9685 Scranton Road, San Diego, California 92121 USA

Received 19 October 2001; accepted 9 March 2002

DOI 10.1002/qua.10417

ABSTRACT: First-principles calculations using the density functional theory code


DMol3 were performed to investigate important pathways in the methanol-to-gasoline
conversion process over zeolite catalysts. Reaction paths and energy barriers involving
the COO bond cleavage and the first COC bond formation were explored using all-
electron periodic supercell calculations and newly implemented algorithms for the
optimization of intermediates and transition states. The simulations indicate that the
formation of surface ylide involves prohibitively high barriers, whereas surface
methoxyl species can easily react with methanol to form ethanol. © 2002 Wiley
Periodicals, Inc. Int J Quantum Chem 91: 467– 473, 2003

formed in the zeolite cage [5, 6]. At that stage di-


Introduction methyl ether can be formed [7], as confirmed by
experiment [1]. However, it is unclear whether di-

T he conversion of methanol to gasoline (MTG)


in zeolites is an important industrial process
that has attracted considerable interest from both
methyl ether is a necessary intermediate for the first
COC bond formation [8]. As for the cleavage of the
COO bond of methanol, it can occur through the
industrial and academic researchers. The main in-
formation of surface methoxyl species [5, 9 –11].
terest is in the mechanism of the first COC bond
This reactive species can then be a starting point for
formation and cleavage of the COO bond of meth-
the formation of the initial COC bond in a reaction
anol. Several reaction mechanisms have been pro-
leading to the formation of ethanol or ethyl-methyl-
posed, and some have supporting experimental ev-
ether [12]. Those intermediates can be easily dehy-
idence [1–3]. A commonly accepted mechanism
drated, thereby producing ethylene, which upon
involves the initial physisorption of methanol at a
Brønsted acid site of the Al-substituted zeolite [4]. further reactions can lead to the formation of higher
As the concentration of methanol increases, clusters olefins, alkanes, etc. [12]. More recently, yet another
of hydrogen-bonded methanol molecules are mechanism for the COC bond formation was pro-
posed: this involves the formation of a surface ylide
Correspondence to: J. Andzelm; e-mail: JWA@ACCELERYS. neighboring a Brønsted acid site [8, 11]. In princi-
COM ple, the surface ylide species could form from the

International Journal of Quantum Chemistry, Vol 91, 467– 473 (2003)


© 2002 Wiley Periodicals, Inc.
ANDZELM ET AL.

methoxyl species by the transfer of a proton to a The Brillouin-zone integration was performed using a
neighboring bridging oxygen. However, it is also 2 ⫻ 2 ⫻ 2 Monkhorst–Pack (MP) grid [18].
possible that the ylide carbon atom is incorporated All geometries were optimized using the re-
into the zeolite framework, leading to a significant cently developed scheme based on delocalized in-
reorganization of a large number of atoms around ternal coordinates generalized to periodic bound-
the Brønsted acid site. ary conditions [16]. It was already demonstrated
Most first-principles computational studies em- that for systems such as zeolites this algorithm can
ploy cluster models to represent the neighborhood be several times more efficient than the one based
of the Brønsted acid site of a zeolite [3, 5, 8 –11]. on Cartesian coordinates. Methanol and water mol-
There are only a few studies using periodic bound- ecules in zeolite pores present additional challenge
ary conditions to represent realistic crystalline en- for the optimization algorithm, because such spe-
vironment for the MTG reaction [4, 6, 7]. The plane cies are typically not connected to the zeolite frame-
wave code CASTEP, for instance, has previously work. However, the DMol3 optimizer allows for a
been used to study the physisorption and clustering treatment of such disconnected fragments without
of methanol molecules leading to the formation of having to introduce any artificial connecting bonds.
dimethyl ether [6, 7]. Here we report the first peri- Several reaction paths investigated here require ap-
odic calculations investigating the formation of a proximate scanning of the potential energy surface.
surface methoxyl species and the formation of an That was accomplished by using internal con-
initial COC bond within a zeolite cage, ultimately straints that were imposed on connected or discon-
leading to ethanol. The possibility of ylide forma- nected fragments. Precise determination of the tran-
tion close to a Brønsted acid site is investigated, as sition states (TS) involving several species in
well. concerted-like reactions in zeolite is obviously a
Density functional theory (DFT) calculations formidable challenge. The traditional method of
were performed using the DMol3 [13, 14] program. calculating a Hessian and following the reaction
Recently developed algorithms to find transition mode can be very costly for any zeolites of practical
states [15] and to optimize structures of a crystal importance. Moreover, the success of such calcula-
[16] were employed to investigate reaction mecha- tions is not guaranteed unless the initial structure is
nisms. The accuracy of DMol3 to predict structures already close to the transition state. In this article
of adsorbed species in zeolites was verified in cal- we employ a recently developed method [15] that
culations on hydrogen-bonded model systems and blends a generalization of the synchronous transit
bridging hydroxyl groups within the zeolite cage. (ST) method of Halgren and Lipscomb with the
conjugate gradient (CG) technique. In this approach
only structures of reactants and products are
needed to locate TS via a series of ST/CG steps.
Computational Details This method was found to be robust, efficient, and
accurate in calculations for many complicated TS of
The DFT calculations reported here were per- reactions in the gas as well as the condensed phase.
formed using the DMol3 program from Accelrys (Ma-
terial Studio 2.0) [13, 14]. The electronic wavefunc-
tions are expanded in atom-centered basis functions
defined on a dense numerical grid. We used the dou-
Results
ble-numeric-polarized (DNP) [13] basis set, and a fine
HYDROGEN BONDING
level of integration grid, amounting to approximately
5500 grid points per atom. The DNP, all-electron basis The mechanism of MTG reaction is determined
set is composed of two numerical functions per va- to a large extent by the network of hydrogen bonds
lence orbital, supplemented with a polarization func- connecting methanol, water molecules, and the H
tion. Each basis function was restricted to within a atom at the Brønsted acid site. A quantitatively
cutoff radius of Rcut ⫽ 4.0 Å, thereby allowing for accurate treatment of hydrogen bonds is, therefore,
efficient calculations without a significant loss of ac- important for this work. To test the accuracy of
curacy. The electron density was approximated using DMol3 in describing H-bond strengths, the interac-
a multipolar expansion up to hexadecapole. Typical tion energies between two methanol molecules and
overbinding associated with local density functionals between two water molecules were computed us-
was rectified through the use of the gradient-cor- ing the PBE/DNP/Rcut ⫽ 4.0 Å settings, as de-
rected Perdew-Burke-Ernzerhof (PBE) functional [17]. scribed in the previous section. The calculated val-

468 VOL. 91, NO. 3


METHANOL CONVERSION

TABLE I ______________________________________________________________________________________________
Relative energies and selected geometrical parameters for the four bridging OH groups in H-Faujasite (FAU).a

␣SiO(H)A1 Energy (kJ/mol)


Occupation site rAlH (Å) (deg) PBE/DNP VWN/DNPb

O1H 2.54 (2.48 ⫾ 0.4) 130.1 (135.7) 0 0


O2H 2.43 141.1 (144.6) 7.8 9.8
O3H 2.46 (2.40 ⫾ 0.4) 136.8 (139.8) 4.1 4.9
O4H 2.44 135.9 (141.9) 9.4 7.9

a
Experimental values are given in parentheses [22, 23].
b
Results of Ref. [23].

ues of 5.8 and 5.7 kcal/mol for methanol and water catalyst, for instance, has 288 atoms. In this article
dimers compare well with experimentally esti- we use the Ferrierite (FER) structure that has eight-
mated ranges of 4.6 –5.9 and 5.0 –5.4 kcal/mol [19], and 10-ring channels—somewhat similar to the 10-
respectively. This good agreement is due to the fact rings channels of ZSM-5. The primitive cell of FER
that the interaction is dominated by the electrostat- contains only 54 atoms and makes the calculations
ics, and DFT predicts dipole moments for both feasible. Recent calculations by Haase and Sauer [4]
methanol and water with reasonable accuracy [20]. on the physisorption of a single methanol molecule
Dispersion forces that are missing in DFT [21] and in the FER and ZSM-5 zeolites show similar com-
contribute perhaps as much as 25% to the total plexes in both zeolites. The methanol adsorption
interaction energy of water dimer [19] are appar- energies in both zeolites differ by less than 4 kcal/
ently compensated by other terms in the interaction mol, which is only 15% of the adsorption energy.
energy. Most results from the calculations on a FER zeolite
The standard procedure in ab initio calculations are therefore expected to be valid for a commercial
of interaction energies uses the so-called basis set zeolite such as ZSM-5.
superposition error (BSSE) approach [19] to account To assess the accuracy of the DMol3 settings in
for incomplete atomic basis sets. The DMol3 pro- a periodic geometry of a zeolite, we performed
gram uses numerical functions that are far more calculations on bridging hydroxyl groups at the
complete than the traditional Gaussian functions,
alumino-silicate Brønsted acid site of Faujasite
and therefore we expect BSSE contribution to be
(FAU) and FER zeolites. Experimental studies
small. We verified this by performing DMol3 calcu-
[22] on the FAU zeolite have revealed that only
lations on a water dimer using a much larger basis
three of the four possible bridging OH groups are
set (five numerical functions per every valence or-
observed, and their relative occupations are as
bital, supplemented with three sets of polarization
follows: 3:1:1.6:0 for O1H:O2H:O3H:O4H sites,
functions, diffuse functions, and Rcut ⫽ 8 Å). This
respectively. In the work by Hill et al. [23] DMol3
calculation affects the interaction energy by less
than 0.6 kcal/mol. Therefore, indirectly, we dem- was applied at the Vosko-Wilk-Nusait (VWN)/
onstrate that any BSSE contributions or further ba- DNP level using unit cell parameters obtained
sis set extensions have a small effect of a fraction of from the shell model. Hill et al. [23] found excel-
kcal/mol on the interaction energy. We are confi- lent correlation between the relative energies of
dent that the PBE/DNP/Rcut ⫽ 4.0 Å DMol3 calcu- the four bridging OH groups and the relative
lations are an accurate and practical approach to experimental occupations. In the current work
study complicated reactions in the unit cell of zeo- full geometry optimization was performed for the
lites. primitive rhombohedral cell with 145 atoms and
fixed cell parameters. Table I summarizes our
results as compared with the results of Ref. [23].
ZEOLITE MODELS AND BRIDGING It is evident that the relative energies from our
HYDROXYL GROUPS
calculations correspond well with the experimen-
Most commercially used zeolite structures are tal site occupations as well as with previous
too large for a detailed study of all reactions in- DMol3 calculations. The bond lengths and bond
volved in the MTG process. The unit cell of a ZSM-5 angles associated with bridging hydroxyl groups

INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 469


ANDZELM ET AL.

2
FIGURE 1. The SN pathway for surface methylation with single methanol. Dotted lines represent hydrogen bonds.
(a) Methanol-hydrogen complex at Brønsted acid site; (b) transition state; (c) surface methoxyl and water hydrogen-
bonded to zeolite.

are within 0.06 Å and 6° of experimental values. group, thereby forming a water molecule [Fig.
2
The O1H site, which is open to the largest channel 1(c)]. This is a strained SN -type reaction with an
of a zeolite, is a preferable location for the hy- activation barrier of 54 kcal/mol. Other DFT or
droxyl group in both FAU and FER zeolites. In ab initio cluster studies [9 –11] have reported a
the case of FER zeolite, the relative energies are 0, barrier of 44 to 65 kcal/mol. Interestingly, the
7.5, 9.5, and 14.2 kJ/mol for the O1H, O2H, O3H, presence of a second methanol molecule lowers
and O4H sites, respectively. The O1H, being the the above activation barrier to 44 kcal/mol, as the
lowest energy site, was selected as the Brønsted TS now corresponds to an unstrained SN 2
pathway
acid site for our MTG study with the FER zeolite. (see Fig. 2). At the TS the surface oxygen, methyl
carbon, and oxygen of the leaving water molecule
FORMATION OF SURFACE METHOXYL are roughly collinear [Fig. 2(b)]. The water mol-
SPECIES ecule is formed as a result of proton transfer to
the methanol hydroxyl group from the methoxo-
A single molecule of methanol forms multiple
nium ion. Our computed barrier of 44 kcal/mol is
hydrogen bonds with the Brønsted acid site of
comparable with the 32– 46 kcal/mol barrier pre-
FER zeolite. There are several adsorption sites
dicted by the MP2/6-31G*//HF/3-21G studies of
possible that are within 1 kcal/mol; Figure 1(a)
Sinclair and Catlow [5] on clusters of various
displays the most stable structure, with an ad-
sorption energy of 18.5 kcal/mol. This value com- sizes.
pares well the experimental estimates of the heat This reaction is facilitated by the hydrogen-
of adsorption in acidic zeolites, ranging from 15 bonded methanol and methoxonium ion [Fig. 2(a)].
to 27 kcal/mol [5, 24]. No protonation of metha- Methoxonium ion is formed spontaneously when a
nol by the Brønsted acid site was found, in agree- methanol molecule captures the proton from the
ment with the recent study by Haase and Sauer Brønsted acid site. No barrier for that reaction is
[4]. The calculated transition state for the meth- found, in agreement with the work by Sandre et al.
ylation of a surface oxygen at the alumino-silicate [7]. Clearly, the presence of the second methanol
Brønsted acid site of FER zeolite is presented in molecule in the zeolite cage facilitates methoxo-
Fig. 1(b). Clearly this is a concerted reaction in- nium ion formation because with one methanol
volving breaking of the COO bond in methanol molecule, no spontaneous deprotonation of zeolite
and bond formation between C and surface oxy- was found. Further analysis of other possible path-
gen. Simultaneously the proton from the Brøn- ways for surface methoxyl formation is given in a
sted acid site is transferred to the hydroxyl separate study.

470 VOL. 91, NO. 3


METHANOL CONVERSION

2
FIGURE 2. The SN pathway for surface methylation with two methanol molecules: (a) methoxonium–methanol com-
plex; (b) transition state; (c) surface methoxyl and methanol–water complex.

REACTIONS OF SURFACE METHOXYL possible pathways, which are discussed in details


SPECIES elsewhere. The most successful pathway involved
Surface methoxyl species can react with metha- water as an important catalytic agent, which
nol or dimethyl ether to form ethanol or methyl- through the network of hydrogen bonds stabilized
ethyl-ether, respectively [12]. Those are the first ethanol molecule and facilitated transfer of the pro-
species containing a COC bond. In this study we ton from the methyl group of the methanol to the
investigated the formation of ethanol. Using the Brønsted acid site of zeolite [Figs. 3(b) and 3(c)].
structure presented in Figure 3(a) as a reactant for The importance of the water molecule in assist-
the COC bond formation, we investigated several ing in this reaction was already noticed by Blasz-

FIGURE 3. Ethanol formation via a methoxyl/water–mediated mechanism. (a) Surface methoxyl and water–methanol
complex; (b) transition state; (c) ethanol, water, and hydrogen bonded to zeolite.

INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 471


ANDZELM ET AL.

FIGURE 4. Formation of the surface ylide species: (a) surface methoxyl species; (b) transition state; (c) surface ylide
species.

kowski and van Santen [12]. The transition state a proton were not successful. The proton always
2
[Fig. 3(b)] indicates that this is an SN -type reaction comes back, without any barrier to form a stable
involving CH3 species already separated from the methoxyl species. However, if there is a rearrange-
zeolite surface with structure close to the trigonal ment of the zeolite framework and the ylide (CH2)
planar. Our computed barrier for that reaction is species is inserted into the AlOO bond, a stable
25.4 kcal/mol, which is significantly less than the 60 intermediate structure is indeed possible. Cluster
kcal/mol barrier reported from the cluster calcula- calculations [8, 25] reveal that the barrier for such
tions of Ref. [12]. The net reaction is exothermic and reaction methoxyl f ylide built in surface is about
the final product adsorbed via hydrogen bonds 50 kcal/mol. We expect a small cluster model to be
[Fig. 3(c)] to the zeolite surface has an energy lower inadequate to study reactions that may significantly
than the reactant by 49.2 kcal/mol. The correspond- affect the entire framework of a zeolite. We inves-
ing reaction energy from Ref. [12] is about 24 kcal/ tigated this reaction (Fig. 4) using a consistent pe-
mol. This disparity can be attributed to differences riodic zeolite model. The calculated barrier of 78.5
in (1) the choice of the model: cluster models were kcal/mol is significantly larger than our previously
used in Ref. [12], as compared with the periodic calculated barriers for methoxyl (44 kcal/mol) and
models used in this work; and (2) the method: the ethanol (25 kcal/mol) formation. Therefore, we
self-consistent GGA calculations were performed can rule out the possibility of a surface ylide for-
here, whereas only perturbative GGA corrections mation. One can notice a significant reorganization
were used in Ref. [12]. From comparing activation of the zeolite framework upon surface ylide forma-
barriers for methoxyl formation we can conclude tion, underlining a need for consistent calculations
that the rate-limiting step is the breaking of the in a periodic environment.
COO bond of methanol and the formation of the
surface methoxyl species [Fig. 2(b)]. The activation
energy for such reaction, 44 kcal/mol, is signifi-
cantly higher than the barrier for the first COC Conclusion
bond formation of 25 kcal/mol [Fig. 3(b)], as found
in this work. Using a DMol3 PBE/DNP approach we per-
Finally, we investigated a recent hypothesis [8, formed calculations on periodic FER zeolite to elu-
25] that the surface ylide (CH2) may provide a cidate the mechanism of key reactions involved in
reactive C atom for the first COC bond formation. the first COO bond breaking and COC bond for-
According to that hypothesis, the methoxyl species mation for the MTG process. This is a consistent
loses a proton, which is transferred to the neighbor- approach that allows us to study significant reor-
ing Brønsted acid site. Our initial attempts to find a ganization of the zeolite cages, which may occur if
stable structure with a surface ylide separated from an adsorbed species is built into the zeolite frame-

472 VOL. 91, NO. 3


METHANOL CONVERSION

work. The main findings of this work are summa- Catlow (Royal Institute of Great Britain, London)
rized below: for valuable discussions. We thank our colleagues
at Accelrys for help in this study and acknowledge
▪ The current approach allows us to calculate members of Accelrys’s Catalysis 2000 Consortium
hydrogen bonding with a reasonable accuracy for their support of this work.
of ⬃0.5 kcal/mol for interaction energies.
▪ Periodic DFT calculations can describe well
the neighborhood of a Brønsted acid site in
alumino-silicate Faujasite zeolite. The experi- References
mentally known positions of the bridging hy-
droxyl groups and their relative stabilities are 1. Chang, C. D.; Silvestri, A. J. J Catal 1977, 47, 249.
correctly reproduced in our calculations. 2. Sauer, J. Chem Rev 1989, 89, 199.
3. Hutchings, G. J.; Hunter, R. Catal Today 1990, 6, 279.
▪ Single methanol molecule is adsorbed in zeo-
4. Haase, F.; Sauer, J. Micro Meso Mat 2000, 35, 379.
lite cages via hydrogen bonds, whereas the
5. Sinclair, P. E.; Catlow, C. R. A. J Chem Soc Faraday Trans
presence of the two methanol molecules al-
1996, 92, 2099.
lows for spontaneous formation of the meth-
6. Shah, R.; Gale, J. D.; Payne, M. C. J Phys Chem 1996, 100,
oxonium ion. 11688.
▪ Formation of surface methoxyl species occurs 7. Sandre, E.; Payne, M. C.; Gale, J. D. Chem Commun 1998, 22,
2
in SN -type concerted reaction with a barrier of 2445.
44 kcal/mol if two methanol molecules are 8. Hutchings, G. J.; Watson, G. W.; Willock, D. J. Micro Meso
present. Mat 1999, 29, 67.
9. Blaszkowski, S. R.; van Santen, R. A. J Phys Chem 1995, 99,
▪ Surface methoxyl species can undergo a reac-
11728.
tion to form an ethanol molecule. The barrier
10. Zicovich-Wilson, C. M.; Viruela, P.; Corma, A. J Phys Chem
for such a reaction is 25 kcal/mol if the water 1995, 99, 13224.
molecule is present. 11. Sinclair, P. E.; Catlow, C. R. A. J Chem Soc Faraday Trans
▪ Formation of the ylide that is built into the 1997, 93, 333.
zeolite cage surface has to be ruled out as the 12. Blaszkowski, S. R.; van Santen, R. A. J Am Chem Soc 1997,
barrier for such a reaction exceeds 78 kcal/ 119, 5020.
mol. 13. Delley, B. J Chem Phys 1990, 92, 508; J Phys Chem 1996, 100,
6107; J Chem Phys 2000, 113, 7756.
▪ The synchronous transit coupled with conju-
14. DMol3, Materials Studio version 2.0 from Accelrys: http://
gated gradient refinement to optimize transi- www.accelrys.com/mstudio/dmol3.html.
tion states [15] and delocalized internals to 15. Govind, N.; Fitzgerald, G.; King-Smith, D. to be published.
optimize structures of reaction intermediates 16. Andzelm, J.; King-Smith, D.; Fitzgerald, G. Chem Phys Lett
[16], under periodic boundary conditions and 2001, 335, 321, and work that is to be published.
internal constraints, allowed fast and accurate 17. Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys Rev Lett 1996,
exploration of reaction pathways and associ- 77, 3865.
ated heats and barriers. 18. Monkhorst, H. J.; Pack, J. D. Phys Rev B 1976, 13, 5188.
19. Szalewicz, K.; Jeziorski, B. In Scheiner, S., Ed. Molecular
Further details of the calculations and structures Interactions; John Wiley & Sons: New York, 1997; p. 3.
of transition states will be published elsewhere. The 20. Klamt, A.; Jonas, V.; Burger, T.; Lohrenz, J. J Chem Phys J
next step in our study of MTG process is likely to Phys Chem 1998, 102, 5074.
include investigations on the role of ether and the 21. Kristian, S.; Pulay, P. Chem Phys Lett 1994, 175, 229.
process of dehydration of ethanol to form ethylene. 22. Czjzek, M.; Jobic, H.; Fitch, A.; Vogt, T. J Phys Chem 1992, 96,
1535.
23. Hill, J.-R.; Freeman, C.; Delley, B. J Phys Chem A 1999, 103,
ACKNOWLEDGMENT
3772.
The authors are grateful to Prof. Joachim Sauer 24. Haase, F.; Sauer, J. J Am Chem Soc 1995, 117, 3780.
(Humboldt University, Berlin) and Prof. Richard 25. Sinclair, P. E.; Catlow, C. R. A. J Phys Chem 1997, 101, 295.

INTERNATIONAL JOURNAL OF QUANTUM CHEMISTRY 473

Vous aimerez peut-être aussi