You are on page 1of 10

Catalysis

Science &
Technology
View Article Online
PAPER View Journal

Kinetics and thermodynamics of polymethylbenzene


formation over zeolites with different pore sizes
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

Cite this: DOI: 10.1039/c6cy00465b


for understanding the mechanisms of methanol to
olefin conversion – a computational study†
Yan-Yan Chen,a Zhihong Wei,a Sen Wang,ab Junfen Li,a Mei Dong,a Zhangfeng Qin,a
Jianguo Wang,a Haijun Jiao*ac and Weibin Fan*a

On the basis of density functional theory including dispersion correction (ωB97XD), the thermodynamics
and kinetics of the formation of polymethylbenzene intermediates in methanol to olefin conversion over
zeolites with different pore sizes have been systematically computed. The agreement between the experi-
mental and theoretical adsorption enthalpies of the several polymethylbenzenes over H-FAU reasonably
validates the applied models and methods, and reveals the importance of dispersion correction in the
space confinement and electrostatic stabilization of the zeolite framework. The free energies of the step-
wise formation of the polymethylbenzenes show that the most favorable active hydrocarbon pool interme-
diates are pentamethylbenzene and hexamethylbenzene over H-BEA and H-SAPO-34, as well as tetra-
methylbenzene over H-ZSM-5 and H-ZSM-22. These stable polymethylbenzenes are also precursors for
the formation of geminal methylated cationic intermediates on the basis of kinetic and thermodynamic
Received 1st March 2016, analyses. The agreement of the thermodynamic and kinetic results on the favorable intermediates validates
Accepted 13th April 2016
the use of Gibbs free reaction energies to estimate the primary component of the intermediates in the var-
ious zeolites. All these pore-size-dependent differences among the zeolites show their enhanced confine-
DOI: 10.1039/c6cy00465b
ment effect, which is mainly influenced by the short-range electrostatic potential including stabilization
www.rsc.org/catalysis and repulsion.

1. Introduction co-catalysts. One of the most important active HCP species is


polymethylbenzenes (polyMBs).15,16
Methanol can be expediently produced via synthesis gas from
multifarious carbon sources such as coal, natural gas and
biomass,1–5 and the conversion of methanol to olefins (MTO)
over acidic zeolite catalysts has been turning into a gradually
important alternative to naphtha cracking for obtaining light
olefins.6–8 In the past few decades, a great deal of effort has
been devoted to elucidate the mechanisms of methanol to
olefin conversion on the basis of experimental and theoretical
studies.9–14 The hydrocarbon pool (HCP) mechanism pro-
posed by Dahl and Kolboe6–8 has received wide recognition,
where it is assumed that hydrocarbon species trapped in zeo-
lite pores interplay with the inorganic framework and serve as

a
State Key Laboratory of Coal Conversion, Institute of Coal Chemistry, Chinese
Academy of Sciences, Taiyuan, Shanxi 030001, China.
E-mail: haijun.jiao@catalysis.de, fanwb@sxicc.ac.cn
b
University of Chinese Academy of Sciences, Beijing 100049, China
c
Leibniz-Institut für Katalyse e.V. an der Universität Rostock, Albert-Einstein-
Strasse 29a, 18059 Rostock, Germany Scheme 1 Schematic representation of the paring and side-chain re-
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ action mechanisms in MTO catalysis (Z–H and Z− stand for the proton-
c6cy00465b ated and deprotonated form of zeolite, respectively).29

This journal is © The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

For polyMBs as active HCP species, two reaction routes, a scopy were observed directly under working conditions using
side-chain route17–20 and a paring route,21 have been proposed only methanol as the reactant. The precursors of these
for the formation of light olefins (Scheme 1). Both routes start carbenium ion intermediates are HCP species of polyMBs
with repeated methylation until the formation of a geminal (for aromatic cycle as an example) which are relatively stable
methylated cation (e.g.; 1,2,3,4,4,5,6-heptamethylcyclohexa-2,5- and easily observable, and some HCP species over a few zeo-
dien-1-ylium ion, 7MB+). Subsequently, the side-chain route lites were identified using isotopic switch experiments.40 On
starts with the formation of an exocyclic double bond by de- H-ZSM-5, lower polyMBs, especially for tetramethylbenzene
protonation, followed by successive side-chain methylation (4MB), were found to act as HCP species.41,42 The main reac-
and side-chain elimination, resulting in olefins. The alterna- tion centers were identified as pentamethylbenzene (5MB)
tive paring route involves contraction and expansion steps and hexamethylbenzene (6MB) in the wide-cavity of H-SAPO-
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

with the elimination of propene. The difference between both 34 (ref. 43 and 44) and in the wide-channel of H-BEA.19,22,36,41
routes is that the paring reaction involves the use of a ring The confinement effect of zeolite frameworks, which re-
carbon to grow an alkyl chain, while in the side-chain reaction, flects the attractive or repulsive interaction between intermedi-
the six-membered ring is conserved during the reaction. ates and frameworks, is decisive for the specific intermediates.
Haw et al.,19 found the side-chain route to be predominant Although some knowledge has been gained about the
in the production of olefins through direct pulse experiments HCP species and MTO reaction mechanism over various
of different polyMBs and 13C-methanol in a large H-BEA pore zeolites,45–51 quantitative analyses of the confinement effect
at 723 K. However, Bjørgen et al.,22 reported the rearrangement on the stability and activity of HCP species in zeolites are not
of 7MB+ by 12C-benzene and 13C-methanol co-reacted at tem- carried out. The confinement effect includes analysis of ther-
peratures below 573 K to be the major reaction route for olefin modynamics about reaction energies and kinetics about activa-
formation on a H-BEA zeolite. These results are reconcilable if tion energies. Investigating the activity of HCP species, the
a change in the dealkylation mechanism occurs with tempera- probability and favorability of the formation of these species
ture.23 Another possible route for light olefin formation, origi- in zeolites should be considered at first by thermodynamics.
nally proposed by Dessau on the basis of successive methyla- Smit et al.52 examined the confinement effect in zeolite catalysis
tion and cracking reactions of C3+ alkenes, has already and argued that simple thermodynamics of molecules adsorbed
attracted extensive attention.24 In more detailed studies using inside zeolite pores can explain the selective formation of prod-
13
C labeling, a complete dual cycle mechanism in a H-ZSM-5 ucts and guide the identification of zeolite structures which
zeolite was proposed, and both the polyMBs cycle and alkene are particularly suitable for desired catalytic applications.
cycle can produce light olefins.25,26 This concept has been fur- In our work, thermodynamic and kinetic analyses on the ba-
ther corroborated by theoretical studies on H-ZSM-5.27,28 sis of density functional theory computations including disper-
MTO mechanisms are determined by intermediates or sive interactions are carried out to quantify the stability and ac-
HCP species which are further controlled by the structure to- tivity of HCP species in various zeolite topologies of H-SAPO-34,
pology of zeolites. Identifying intermediates is of great impor- H-ZSM-5, H-BEA and H-ZSM-22, and the confinement effect of
tance to understand reaction mechanisms as well as the the zeolite frameworks on the HCP species is also discussed.
structure and activity correlation of a reaction. Although For clear presentation, only the HCP species of polyMBs are
many experimental techniques have been used to detect HCP considered in this work. Kinetic analyses about the gem-
intermediates, and some intermediates over only a few zeo- methylation of polyMBs on H-CHA, H-ZSM-5 and H-BEA with-
lites have been found, specific HCP intermediates over differ- out dispersion correction were reported by Waroquier et al.,53
ent zeolites are less known. Some carbenium ions in the and a similar protocol for thermodynamic analyses on free en-
MTO reaction, such as 5MB+, cyclopentenyl cation and 7MB+, ergies of methylbenzenes was recently reported by Wang et al.54
were identified by using in situ solid-state NMR, whereas the In contrast to the above two references, we also included the
formation of these carbenium ions in catalysts was verified interconversion of different MBs proceeding through a methyla-
mostly indirectly or by the aid of the reaction of co-feeding tion step in our calculations. The advantages of our work are:
methanol and benzene or methylbenzene.20,30–33 Dai et al.,34 (a) identification of more favorable reactions as well as active
observed several initial species (i.e., three-membered ring hydrocarbon intermediates by comparing the energy barriers of
compounds, dienes, polymethylcyclopentenyl, and poly- the competitive reactions of methylation and gem-methylation
methylcyclohexenyl cations) on a H-SAPO-34 model catalyst of polyMBs; (b) the validity of the Gibbs free reaction energies
during the early stages of MTO conversion via 1H MAS NMR to estimate the primary component of intermediates in zeolites
and 13C MAS NMR, and proposed an olefin-based catalytic cy- on the basis of the agreement between thermodynamic and ki-
cle as the primary reaction pathway before the steady-state, netic results; (c) the essence of the confinement effect analyzed
with highly active polymethylbenzenium ions as the most im- by comparing the energies in the four zeolites.
portant intermediates. Until now, only cyclopentenyl cations
and 5MB+ ions on H-ZSM-5 (ref. 35 and 36) and cyclopentenyl 2. Computational model and methods
cations on H-ZSM-5 by applying in situ ultraviolet-visible
spectroscopy,37 and 7MB+ ions on DNL-6 (ref. 38) and CHA To model zeolite catalysts, it is essential to choose appropri-
(H-SAPO-34 and H-SSZ-13)39 by applying MAS NMR spectro- ate systems, taking into account both the space confinement

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

and electrostatic effects of the framework. Principally, peri- along the pre-existing Si–O bonds, and the terminal Si–H
odic models taking an entire unit cell into account should be (Al–H and P–H for H-SAPO-34) distances are 1.47 Å (1.55 and
the most suitable in capturing the true nature of nano- 1.35 Å for H-SAPO-34). We used these cluster models
porous materials. Indeed, such models were utilized to inves- containing a 5T [6T for H-SAPO-34] active region for
tigate the mechanism and kinetics of MTO processes.55,56 SiOHAlIJOSi)2OSi [(PO)2SiOHAl(OSi)2 for H-SAPO-34] along
Contrary to periodic models, small clusters might not reason- with the adsorbates in the channels to be relaxed, while the
ably represent real structures, but clusters large enough can rest of the structure is fixed at their crystallographic coordi-
be used to account for the distinguishing topological features nates during geometry optimization.
of zeolite catalysts,57–59 and to rationally describe the space In this work, all calculations were done using large clus-
confinement and electrostatic stabilization effects of zeolite ters with the Gaussian 09 package.67 All the transition states
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

frameworks. (TS) were guessed by using the OPT = TS method and con-
In our work, 52T H-BEA, 49T H-ZSM-5, 61T H-FAU, 45T H- firmed by the quasi-internal reaction coordinate approach to
ZSM-22 and 46T H-SAPO-34 cluster models were used verify each authentic transition state bridging the corre-
(Fig. 1). The initial structures of these zeolites were taken sponding reactants and products. Furthermore, the transition
from the database of the Structure Commission of the Inter- state is a first-order saddle point on the potential energy sur-
national Zeolite Association (IZA).60,61 The H-SAPO-34 struc- face with only a single imaginary frequency, and the obtained
ture consists of a super-cage (the maximum diameter of a reactants and products were verified to be situated in the en-
sphere that can be included is 7.4 Å) made up of small pores ergy minima points of the potential energy surface with only
(3.8 × 3.8 Å), which is derived from a silicalite-CHA structure, real frequencies. The B3LYP functional and the 6-31G(d,p)
where all the symmetrically equivalent Si atoms are basis set are used in all the geometry optimizations and fre-
substituted by P and Al atoms in an alternating manner, and quency calculations.
one P atom is replaced by one Si atom to generate one To obtain accurate interaction schemes, single-point en-
Brønsted acid site per cage.62 The H-BEA structure consists of ergy was calculated using the ωB97X-D functional including
three-dimensional large intersecting straight channels (5.5 × dispersion corrections on the B3LYP/6-31G(d,p) optimized ge-
6.7 Å), with an Al atom located at T9,63 and the maximum di- ometries. The ωB97X-D functional is promising for main
ameter of a sphere that can be included is 6.68 Å. The H- group thermochemistry, kinetics and non-covalent interac-
ZSM-5 structure consists of an intersecting channel between tion;68,69 and the 6-311+G(2df,2p) basis set was used. The
a straight channel (5.3 × 5.6 Å) and a zigzag channel (5.1 × ωB97X-D functional was recently found to perform very well
5.5 Å), with an Al atom located at T12 (ref. 64) and the maxi- for the adsorption and reactions on zeolites.70
mum diameter of a sphere that can be included is 6.4 Å. The The dispersion corrected adsorption energy (Eads-d) was
H-ZSM-22 structure consists of a one-dimensional straight calculated using eqn (1), where Etot-d(HCP/Z) is the dispersion
channel (5.7 × 4.6 Å) with an Al atom located at T4,65 and the corrected total energy for a zeolite with absorbed species in
maximum diameter of a sphere that can be included is 5.7 Å. its equilibrium geometry, Etot-d(Z) is the dispersion corrected
The H-FAU structure consists of super-cages that are inter- total energy of a zeolite, and Etot-d(HCP) is the dispersion
connected by 12-membered rings with dimensions of 7.4 × corrected total energy of the free hydrocarbon species in the
7.4 Å; the same as H-SAPO-34, all its tetrahedral Si atoms are gas phase. The adsorption energy without dispersion correc-
symmetrically equivalent, and the maximum diameter of a tion (Eads) was calculated using a similar method as in
sphere that can be included is 11.2 Å.66 All the peripheral Si eqn (1), but the Etot value used is the total energy using
(Al and P for H-SAPO-34) atoms were saturated by H atoms B3LYP/6-31G(d,p). The adsorption energy without dispersion

Fig. 1 H-SAPO-34-46T, H-BEA-52T, H-ZSM-22-45T, H-ZSM-5-49T, H-FAU-61T cluster models.

This journal is © The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

correction (Eads') using ωB97X/6-311+G(2df,2p) was also calcu- −1.30 eV, respectively). Over H-BEA, 5MB has the strongest
lated, and the results are shown in Table S1.† adsorption enthalpy (−1.37 eV). Over H-ZSM-5, 1,3,5-3MB and
1,2,3,5-4MB have the strongest adsorption enthalpies (−1.42
Eads‐d = Etot‐d(HCP/Z) − Etot‐d(HCP) − Etot‐d(Z) (1) and −1.41 eV, respectively). Over H-ZSM-22, both p- and
m-xylenes have the strongest adsorption enthalpies (−1.32
and −1.37 eV, respectively), while 6MB has a very positive ad-
3. Results and discussion sorption energy (1.56 eV), indicating its considerable repul-
sive interaction. It is interesting to note that the strongest ad-
3.1. Adsorption enthalpies of polyMBs sorption enthalpies of polyMBs over the four zeolites are
The initial configurations of each polyMB in the zeolite pores within 1.33–1.42 eV.
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

are multitudinous, but their energetic differences are small


(within 0.2 eV). The optimized configurations with the stron- 3.2. Free energies of polyMB formation
gest adsorption enthalpies are used in this work. Table 1 lists To identify the formed polyMB intermediates in the zeolite
the adsorption enthalpies of polyMBs at 525 K. On the basis pores, it is crucial to consider their Gibbs free energy (ΔRGn)
of the adsorption enthalpies of some polyMBs in a H-FAU ze- of formation on the basis of the reaction eqn (2), where
olite calculated through experiment at 525 K,71 we validated methanol is the only starting material and interacts with zeo-
our method on the basis of a 61T-FAU cluster model at first. lite (ZOH) leading to the formation of polyMBs, and n can be
As given in Table 1, the dispersion corrected adsorption en- considered as the number of the methyl groups on the ben-
thalpies agree well with the experimental data and the corre- zene ring. To balance the mass, we simply used molecular H2
lation coefficient (R 2) is 0.895, while those without dispersion as the gas product; however, the contribution of H2 is a con-
correction are much smaller. This indicates that it is crucial stant and does not change with the change in the methyl
to include dispersion correction for calculating the adsorp- groups. This will not affect the relative energies and also does
tion of polyMBs in the zeolites. It also indicates that disper- not affect our results and conclusion. The computed ΔRGn
sion correction depends on the size of the adsorbed mole- value allows the discrimination of the preferred intermediate,
cules; the larger the adsorbed molecule, the larger the i.e., the more negative the ΔRGn value, the more preferred the
correction, e.g. the dispersion correction is 0.45 eV for ben- formed intermediate. Table 2 shows the computed free ener-
zene, while 1.07 eV for 6MB. Benchmark studies with even gies of formation at 725 K, and this is because the typical
larger cluster models and different methods show that our MTO reaction conditions are 625–725 K. The temperature
current model is sufficient enough in reproducing the ad- effect at 525 and 725 K on the stability of the polymethyl-
sorption enthalpy of hexamethylbenzene in the H-FAU zeolite benzenes is calculated and analyzed. The same trend is
in comparison with the experimental data. It is also found found, and the gap is basically identical (Fig. S1†). Thus, the
that increasing the size of the model for H-ZSM-5 does not af- result at 725 K is used in this work.
fect the adsorption enthalpy of hexamethylbenzene signifi-
cantly (Table S2†). Therefore, we used our current models in ZOH + (6 + n)CH3OH(g) = ZOH − C6H6−n(CH3)n
discussing the adsorption enthalpies of polymethylbenzene + (6 + n)H2O(g) + 3H2(g) (2)
over different zeolites.
Based on the agreement between the theoretical and ex- Within the four zeolites, the free reaction energies of ben-
perimental results over H-FAU, we computed the adsorption zene formation are very close, indicating that benzene is a
enthalpies over other zeolites. Over H-SAPO-34, 1,2,4,5-4MB rather small molecule, and its formation is not significantly
and 5MB have the strongest adsorption enthalpies (−1.33 and affected by the framework of these zeolites. The same results

Table 1 Adsorption enthalpies (eV) with (Hads-d) and without (Hads; in parenthesis) dispersion correction at 525 K

PolyMBs H-SAPO-34 H-BEA H-ZSM-5 H-ZSM-22 H-FAU H-FAU (expt)71


Benzene (B) −0.77 (−0.16) −0.80 (−0.20) −0.88 (−0.22) −1.0 (−0.07) −0.74 (−0.29) −0.78
Toluene (MB) −0.91 (−0.15) −0.94 (−0.24) −1.06 (−0.23) −1.16 (−0.05) −0.88 (−0.31) −0.85
p-Xylene (p-2MB) −1.06 (−0.12) −1.09 (−0.27) −1.20 (−0.17) −1.32 (−0.06) −0.97 (−0.35) −0.93
m-Xylene (m-2MB) −1.07 (−0.13) −1.12 (−0.25) −1.26 (−0.21) −1.37 (−0.06) −0.98 (−0.35)
o-Xylene (o-2MB) −1.0 (−0.08) −1.07 (−0.27) −1.17 (0.01) −1.03 (0.35) −0.99 (−0.34)
Mesitylene (1,3,5-3MB) −1.17 (0.07) −1.13 (−0.23) −1.42 (−0.19) −0.79 (0.91) −1.06 (−0.36) −1.01
1,2,4-Trimethylbenzene (1,2,4-3MB) −1.19 (−0.04) −1.16 (−0.23) −1.23 (−0.01) −1.20 (0.34) −1.06 (−0.35)
1,2,3,5-Tetramethylbenzene (1,2,3,5-4MB) −1.07 (0.3) −1.22 (−0.18) −1.41 (0.11) −0.31 (1.66) −1.08 (−0.29) −1.17
1,2,4,5-Tetramethylbenzene (1,2,4,5-4MB) −1.33 (−0.02) −1.21 (−0.14) −0.68 (0.92) −0.89 (1.01) −1.10 (−0.28)
1,2,3,4-Tetramethylbenzene (1,2,3,4-4MB) −1.17 (0.22) −1.14 (−0.07) −0.55 (1.08) −0.30 (1.57) −0.98 (−0.25)
Pentamethylbenzene (5MB) −1.30 (0.29) −1.37 (−0.18) −0.81 (1.04) −0.04 (2.19) −1.27 (−0.28)
Hexamethylbenzene (6MB) −0.90 (1.01) −1.27 (0.13) −0.09 (2.35) 1.56 (4.26) −1.24 (−0.17) −1.30
Maximum diameter (Å)a 7.4 6.7 6.4 5.7 11.2
a
http://www.iza-structure.org/databases.

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

Table 2 Computed Gibbs free energies of the formation reaction of polyMB (ΔRGn, eV, based on eqn (2)) and their sequential difference between (n +
1)MB and nMB (ΔΔRG, eV, based on eqn (3), in square brackets)

Hydrocarbons H-SAPO-34 H-BEA H-ZSM-5 H-ZSM-22 H-FAU


B −4.02 [0.00] −4.07 [0.00] −4.16 [0.00] −4.08 [0.00] −4.07 [0.00]
MB −4.89 [−0.87] −4.82 [−0.75] −4.89 [−0.73] −4.85 [−0.77] −4.85 [−0.78]
p-2MB −5.69 [−0.80] −5.58 [−0.76] −5.59 [−0.70] −5.70 [−0.85] −5.54 [−0.69]
m-2MB −5.51 [−0.62] −5.44 [−0.62] −5.71 [−0.82] −5.71 [−0.86] −5.57 [−0.77]
o-2MB −5.63 [−0.74] −5.35 [−0.53] −5.22 [−0.33] −5.12 [−0.27] −5.47 [−0.67]
1,3,5-3MB −6.02 [−0.51] −6.09 [−0.65] −6.43 [−0.72] −5.48 [+0.23] −6.33 [−0.76]
1,2,4-3MB −6.34 [−0.65a] −6.31 [−0.73] −6.26 [−0.67] −5.94 [−0.24] −6.24 [−0.70]
1,2,3,5-4MB −6.78 [−0.44] −7.02 [−0.71] −6.93 [−0.67] −5.57 [+0.37] −6.89 [−0.65]
−7.03 −6.97 [−0.66] −6.03 [+0.23] −6.16 −6.77
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

1,2,4,5-4MB [−0.69] [−0.22] [−0.53]


1,2,3,4-4MB −6.77 [−0.43] −6.86 [−0.55] −5.78 [+0.48] −5.38 [+0.56] −6.64 [−0.40]
5MB −7.22 [−0.44; −0.19b] −7.58 [−0.56; −0.61b] −6.43 [+0.50] −5.53 [+0.63] −7.28 [−0.51]
6MB −7.06 [+0.16] −7.78 [−0.20] −5.87 [+0.56] −3.97 [+1.56] −7.73 [−0.45]
a
−0.65 eV is the difference value between p-2MB and 1,2,4-3MB. b
−0.44 eV is the difference value between 1,2,3,5-4MB and 5MB; −0.19 eV is
the difference value between 1,2,4,5-4MB and 5MB.

are also found for toluene, where the largest difference is shown in eqn (3). Indeed, such sequential Gibbs free energy
only 0.07 eV. change can reveal the methylation probability, i.e., a negative
From the xylene isomers, shape dependence on the cage ΔΔRG value implies the formation of (n + 1)MB to be thermo-
size of the zeolites is observed, especially for H-ZSM-5 and dynamically possible and favorable starting from nMB, a pos-
H-ZSM-22. H-ZSM-5 favors m-xylene (−5.71 eV), by 0.49 eV itive ΔΔRG value reveals the formation of (n + 1)MB to be
higher than o-xylene (−5.22 eV); H-ZSM-22 favors both m- and thermodynamically less or not possible in the zeolites, and a
p-xylenes (−5.71 and −5.70 eV), by 0.59 and 0.58 eV higher ΔΔRG value close zero reveals the possible co-existence and
than o-xylene (−5.12 eV); H-SAPO-34 and H-BEA favor p-xylene equilibrium of (n + 1)MB and nMB in a zeolite.
(−5.69 and −5.58 eV, respectively); while three xylene isomers
show similar free reaction energies in H-FAU. Although the ZOH − C6H6−n(CH3)n + CH3OH(g) = ZOH − C6H6−n−1(CH3)n+1
most favorable xylenes are not the same isomer, their free en- + H2O(g) (3)
ergies are very close, which indicates that their formation is
not significantly affected by the framework of these zeolites. Within the four zeolites, the formation of toluene from
For both 3MB isomers, H-ZSM-5 and H-FAU favors 1,3,5- benzene and the formation of xylenes from toluene are ther-
3MB (−6.43 and −6.33 eV), while 1,2,4-3MB is more favored modynamically favorable (Table 2); therefore, the transforma-
over H-SAPO-34 (−6.34 eV), H-BEA (−6.31 eV) and H-ZSM-22 tion from benzene via toluene to xylenes should be easy and
(−5.94 eV). It is found that the free energy of 3MB over they do not represent stable intermediates in the zeolites.
H-ZSM-22 is lower than those of the other zeolites, indicating Starting from toluene, the more favorable methylation routes
the larger repulsive interaction between the framework and are shown in Scheme 2.
3MB. The same results are also found for the 4MB isomers. In H-SAPO-34, both p-2MB and o-2MB isomers are more
H-ZSM-5 favors 1,2,3,5-4MB (−6.93 eV), by 0.90 eV higher favorable from toluene [ΔΔRG = −0.80 and −0.74 eV, respec-
than 1,2,4,5-4MB (−6.03 eV), while H-SAPO-34 (−7.03 eV) and tively], and they can be transformed into the next more favor-
H-ZSM-22 (−6.16 eV) favor 1,2,4,5-4MB, and H-BEA as well as able 1,2,4-3MB [ΔΔRG = −0.65 and −0.71 eV, respectively]. The
H-FAU favors both 4MB isomers. next more favorable intermediate is 1,2,4,5-4MB starting from
With increasing size, higher polyMBs experience strong re- 1,2,4-3MB [ΔΔRG = −0.69 eV]. Considering the small energy
pulsive interactions counterbalancing vdW stabilization. For changes from 1,2,4,5-4MB to 5MB (−0.19 eV), and from 5MB
5MB, H-BEA has the largest free energy (−7.58 eV), followed to 6MB (+0.16 eV), all three intermediates might be formed
by H-FAU (−7.28 eV) and H-SAPO-34 (−7.22 eV), while those of in equilibrium and 5MB should be the major isomer.
H-ZSM-5 (−6.43 eV) and H-ZSM-22 (−5.53 eV) are much lower. In H-BEA, the most favored methylation route for the for-
For 6MB, H-BEA has the largest free energy (−7.78 eV), mation of polyMBs is p-2MB → 1,2,4-3MB → 1,2,3,5-4MB and
followed by SAPO-34 (−7.06 eV), while those of H-ZSM-5 1,2,4,5-4MB; 5MB and 6MB might be formed in equilibrium,
(−5.87 eV) and H-ZSM-22 (−3.97 eV) are much lower. and 6MB should be the major isomer, since the formation of
All the computed Gibbs free energies of the formation re- 6MB is more favorable by 0.20 eV.
actions of polyMBs show their thermodynamic probability. In H-ZSM-5, the most favored methylation route for the
Since the formation of polyMBs is considered as sequential formation of polyMBs is m-2MB → 1,3,5-3MB → 1,2,3,5-4MB.
methylation reactions (Scheme 1), we become interested in However, the formation of 5MB and 6MB is endergonic by
the sequential Gibbs free energy change [ΔΔRG = ΔRGn+1 − 0.50 and 0.56 eV, respectively, and therefore not accessible.
ΔRGn] from ZOH-C6H6−n(CH3)n to ZOH-C6H6−n−1(CH3)n+1, as The most favorable intermediate should be 1,2,3,5-4MB.

This journal is © The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology


Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

Scheme 2 Most favored methylation route for the formation of polyMBs on the basis of the reaction free energies (eV) in eqn (2) on different
zeolites.

In H-ZSM-22, the most favored methylation route for the and gem-methylation are parallel and competitive, their selec-
formation of polyMBs is p/m-2MB → 1,2,4-3MB → 1,2,4,5- tivity can be qualitatively estimated by their kinetic accessibil-
4MB. However, the formation of 5MB and 6MB is endergonic ity and thermodynamic stability, i.e., the lower the barrier
by 0.63 and 1.56 eV, respectively, and not accessible. There- and the stronger the formation energy, the more selective the
fore, the most favorable intermediate should be 1,2,4,5-4MB reaction. Benchmark calculations for the free energy barrier
at equilibrium. and reaction enthalpy for the reaction from 4MB to 5MB over
In H-FAU, the situation is very clear and simple, since all H-ZSM-5 by using larger cluster models and different
the methylation steps are thermodynamically favorable methods validated our current model (Table S3†). In addi-
(Table 2). The large Gibbs free energy (−0.45 eV) of formation tion, we also compared our current free energy barriers for
of 6MB shows that 6MB is the most favorable intermediate at the stepwise methylation and gem-methylation over different
equilibrium, and there is much less confinement of the cage zeolites with available data in the literature (Table S4†). In
on the one hand. On the other hand, the formed 6MB should several cases, large differences are found, and this is mainly
be further transformed into polycyclic aromatic hydrocarbons. due to the differences in models and methods, such as func-
The computed free reaction energies show good correla- tional, basis set, dispersion correction, thermal correction
tion between the polyMB probability and the cage size. For and entropy contribution, as well in reaction paths such as
example, over the large-sized H-SAPO-34 (7.37 Å) and H-BEA concerted or stepwise. The critical bond distances of the tran-
(6.68 Å), 5MB and 6MB are favorable in equilibrium, while sition states of methylation and gem-methylation are given in
over the medium-sized H-ZSM-5 (6.36 Å) and small-sized Table S5.†
H-ZSM-22 (5.71 Å), 4MB is the favorable intermediate. In H-BEA, the methylation free energy barrier from 2MB
to 3MB as well as from 3MB to 4MB is 1.43 and 1.23 eV, re-
spectively, and they are virtually the same as those of the
3.3. Kinetics of polyMB methylation competitive gem-methylation to 3MB+ and 4MB+ (1.43 and
In addition to the computed thermodynamic probability of 1.31, respectively). However, the methylation from 2MB to
polyMB formation on the basis of reaction free energies 3MB as well as from 3MB to 4MB is much more thermody-
starting from methanol, the free energy barriers and reaction namically favorable (−0.73 and −0.83 eV, respectively) than
free energies of methylation and gem-methylation of polyMBs gem-methylation (0.35 and 0.13 eV, respectively), which indi-
using surface methoxy groups for methylation72–74 are also cates that the reverse reaction of gem-methylation is more fa-
calculated to investigate the activity of HCP species of poly- vorable. The methylation free energy barrier from 4MB to
MBs (Fig. 2 and Table 3). It is noted that methylation of nMB 5MB as well as from 5MB to 6MB is 1.17 and 1.07 eV, respec-
to (n + 1)MB includes geminal methylation and deproton- tively, and they are lower than those of the competitive gem-
ation. In Fig. 2, the deprotonation step is skipped because of methylation to 5MB+ and 6MB+ (1.61 and 1.50 eV, respec-
the much lower energy barriers12,56 where the potential en- tively). In addition, the methylation from 4MB to 5MB as well
ergy surface of (n + 1)MB can also be seen. Since methylation as from 5MB to 6MB is more thermodynamically favorable

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

(−0.74 and −0.46 eV, respectively) than gem-methylation (0.04


and −0.06 eV, respectively). The gem-methylation of 6MB to
7MB+ requires a free energy barrier of 1.23 eV and is exer-
gonic by 0.29 eV. Thus, 6MB should be the more active spe-
cies in the aromatic cycle for the gem-methylation reaction to
form 7MB+. The potential energy surface in Fig. 2 shows that
methylation is more kinetically and thermodynamically favor-
able than competitive gem-methylation from 2MB to 6MB;
6MB is the most favored intermediate and 7MB+ is the most
favored geminal methylated intermediate.
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

In H-SAPO-34, methylation is more kinetically and thermo-


dynamically favorable than competitive gem-methylation from
2MB to 5MB, and the methylation and gem-methylation of
5MB are competitive. Therefore, both 5MB and 6MB should
be the active species in the aromatic cycle for the gem-
methylation reactions to form 6MB+ and 7MB+.
For H-ZSM-5, methylation is more favorable from 2MB to
4MB, but gem-methylation is more kinetically and thermody-
namically favorable than competitive methylation from 4MB
and 5MB. Since 4MB is the most favored intermediate, and
5MB and 6MB are higher in energies and are not kinetically
and thermodynamically accessible, 4MB should be the active
species in the aromatic cycle for the gem-methylation reaction
to form 5MB+, similar to H-ZSM-22.

3.4. Confinement effect on thermodynamic and kinetic


energies
Apart from the analysis of the reaction free energies, the
contributions of entropy (TΔS) at 725 K, electrostatic poten-
tial (ΔEe) and long-range van der Waals dispersion (ΔEdis)
(Table S6†) are also interesting for investigating the confine-
ment effect of the zeolite frameworks on polyMB formation,
where ΔEdis is calculated as the difference of the electronic
energies between ωB97X-D/6-311+G(2df,2p) and ωB97X/6-
311+G(2df,2p).
As shown in Fig. 3, the individual contributions of each
term in the four zeolites are similar for low polyMBs, but dif-
fer to a large extent for high polyMBs. As we know, the dis-
persion correction energy varies directly with C/R 6, where R
represents the distance between a pair of atoms, and C repre-
sents the correction coefficient. Therefore, the shorter the
distance between a pair of atoms, the larger the dispersion
correction energy. Following the decrease in the zeolite chan-
nel size in the order H-BEA > H-SAPO-34 > H-ZSM-5 > H-
ZSM-22, the contributions of ΔEdis increase because of the
larger stabilization by the long-range electrostatic potential
on high polyMBs in the small channel; however, this stabili-
zation conversely leads to a decrease in TΔS, while ΔEe rap-
idly decreases because of the strong steric hindrance between
the framework and high polyMBs in the small channel. For
testing the influence of a relaxed model on steric effect, a
larger 8T relaxed model over H-ZSM-5 was used. The com-
Fig. 2 Potential energy surfaces on the basis of activation free
energies ΔG ≠ and reaction free energies ΔGr at 725 K for polyMB
puted Gibbs free reaction energy of hexamethylbenzene for-
methylation (nMB → (n + 1)MB) and gem-methylation (nMB → (n + 1) mation is lower (−6.05 eV), which corresponds to a change of
MB+) using surface methoxy groups for methylation. only 3%, indicating that the size of the relaxed model shows

This journal is © The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

Table 3 Calculated free energy barrier (ΔG ≠), enthalpy barrier (ΔH ≠, eV) and entropy loss (−TΔS ≠, eV) as well as reaction free energies ([ΔGr], in square
brackets) at 725 K of methylation (nMB → (n + 1)MB) and gem-methylation (nMB → (n + 1)MB+) of polyMBs. Here, 2MB is p-2MB; 3MB is 1,2,4-3MB;
4MB is 1,2,3,5-4MB in H-SAPO-34, H-BEA and H-ZSM-5, while 1,2,4,5-4MB in H-ZSM-22; 3MB+ is a 1,4,4-trimethylcyclohexa-2,5-dien-1-ylium ion;
4MB+ is a 1,3,4,4-tetramethylcyclohexa-2,5-dien-1-ylium ion; 5MB+ is a 1,3,4,4,5-pentamethylcyclohexa-2,5-dien-1-ylium ion in H-SAPO-34, H-BEA and
H-ZSM-5, while a 1,2,4,4,5-pentamethylcyclohexa-2,5-dien-1-ylium ion in H-ZSM-22

H-SAPO-34 H-BEA H-ZSM-5 H-ZSM-22


ΔG ≠ [ΔGr] ΔH ≠ −TΔS ≠ ΔG ≠ [ΔGr] ΔH ≠ −TΔS ≠ ΔG ≠ [ΔGr] ΔH ≠ −TΔS ≠ ΔG ≠ [ΔGr] ΔH ≠ −TΔS ≠
2MB → 3MB 1.71 [−0.65] 1.43 0.28 1.43 [−0.73] 1.18 0.25 1.25 [−0.90] 1.03 0.22 0.92 [−0.95] 0.82 0.10
2MB → 3MB+ 1.71 [0.85] 1.45 0.26 1.43 [0.35] 1.12 0.31 1.23 [0.34] 1.02 0.21 1.15 [0.37] 1.08 0.08
3MB → 4MB 1.68 [−0.36] 1.47 0.19 1.23 [−0.83] 1.06 0.17 1.32 [−0.96] 1.16 0.16 0.84 [−0.84] 0.88 −0.04
3MB → 4MB+ 1.59 [0.38] 1.36 0.23 1.31 [0.13] 0.97 0.34 1.09 [−0.09] 0.92 0.17 1.47 [0.34] 1.37 0.10
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

4MB → 5MB 1.74 [−0.52] 1.52 0.22 1.17 [−0.74] 0.96 0.21 1.59 [0.24] 1.27 0.32 1.34 [−0.13] 1.10 0.24
4MB → 5MB+ 1.84 [0.16] 1.53 0.31 1.61 [0.04] 1.20 0.41 1.17 [−0.27] 1.23 0.30 1.10 [−0.13] 0.92 0.18
5MB → 6MB 1.71 [0.18] 1.49 0.22 1.07 [−0.46] 0.92 0.15 1.39 [0.32] 1.24 0.15 1.60 [−0.17] 1.42 0.18
5MB → 6MB+ 1.81 [0.09] 1.57 0.24 1.50 [−0.06] 1.15 0.35 0.99 [−0.38] 0.86 0.13 0.90 [0.09] 0.83 0.07
6MB → 7MB+ 1.51 [0.14] 1.34 0.17 1.23 [−0.29] 0.93 0.30 1.04 [−0.44] 0.91 0.13 1.02 [−0.06] 0.93 0.09

goal is to evaluate the confinement effect of the framework


on the stability and activity of the specific hydrocarbon pool
species. All the intermediates were optimized and character-
ized at the B3LYP/6-31G(d,p) level. To take dispersion forces
into account, we used the long-range-corrected ωB97XD/6-
311+G(2df,2p) hybrid functional to calculate the energies on
the B3LYP/6-31G(d,p) optimized geometries.
On the basis of the agreement between the experimental
and theoretical adsorption enthalpies of several methyl-
benzenes over H-FAU, it is found that it is very necessary to
include the van der Waals dispersion correction in determin-
ing the space confinement and electrostatic stabilization of
the zeolite framework. Such agreements reasonably validate
our applied cluster models and computational methods.
Without dispersion correction, the computed adsorption en-
Fig. 3 Individual contributions from entropy (TΔS), electronic ergies are pathologically lower than the experimental values.
dispersion energy (ΔEdis) and electrostatic energy (ΔEe) at 725 K. The computed free energy of the methylation reactions
shows the correlation between the polymethylbenzene proba-
bility and the cage size. Over the large-sized H-SAPO-34 and
a little influence on steric effect, and the current model sys- H-BEA, pentamethylbenzene and hexamethylbenzene are favor-
tems are reasonable to describe the formation of polyMBs in able in equilibrium, while over the medium-sized H-ZSM-5 and
zeolites. In addition, with the size increase of polyMBs, the small-sized H-ZSM-22, tetramethylbenzene is the favorable
degree of individual energies varied as follows: ΔEe > TΔS > intermediate. These stable polymethylbenzenes are the precur-
ΔEdis. sors for the formation of geminal methylated cationic interme-
Apart from the analysis of activated energies, the enthalpy diates on the basis of kinetic and thermodynamic analyses.
barriers and entropy barriers of polyMB methylation are also The same trend of thermodynamics and kinetics on the
analyzed (Table 3). In H-SAPO-34, the intrinsic enthalpy bar- favorable intermediates validates the applicability of Gibbs
riers are higher than those of other zeolites owing to the in- free reaction energies to estimate the primary component of
sufficient stabilization of transition states in the large cage the intermediates in the various zeolites. All these differences
pores. Comparing the enthalpy barriers, the entropy barriers among these zeolites with different pore sizes show their
in the four zeolites are small. enhanced confinement effects, which is mainly influenced by
the short-range electrostatic potential including stabilization
4. Conclusions and repulsion.

The thermodynamics and kinetics of the formation of poly- Acknowledgements


methylbenzene intermediates over H-SAPO-34, H-BEA, H-
ZSM-5 and H-ZSM-22 with different shapes and pore sizes are We are grateful for the financial support from the National
investigated on the basis of density functional theory includ- Basic Research Program (2011CB201406), the National Natu-
ing van der Waals dispersive corrections (ωB97XD). Our ral Science Foundation of China (21573270, 21103216,

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

21273263 and 21273264), the Natural Science Foundation of 27 D. Lesthaeghe, J. Van der Mynsbrugge, M. Vandichel, M.
Shanxi Province of China (2013021007-3), the Coal Based Low Waroquier and V. Van Speybroeck, ChemCatChem, 2011, 3,
Carbon Joint Foundation of NSFC and Shanxi Province of 208–212.
China, U1510104) and the CAS/SAFEA International Partner- 28 C. M. Wang, Y. D. Wang and Z. K. Xie, J. Catal., 2013, 301,
ship Program for Creative Research Teams. The computa- 8–19.
tional results were obtained from the ScGrid of the Super- 29 D. Lesthaeghe, A. Horré, M. Waroquier, G. B. Marin and V.
computing Center, Computer Network Information Center of Van Speybroeck, Chem. – Eur. J., 2009, 15, 10803–10808.
Chinese Academy of Sciences. 30 T. Xu and J. F. Haw, J. Am. Chem. Soc., 1994, 116, 7753–7759.
31 T. Xu, D. H. Barich, P. W. Goguen, W. G. Song, Z. K. Wang,
References J. B. Nicholas and J. F. Haw, J. Am. Chem. Soc., 1998, 120,
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

4025–4026.
1 D. A. Hickman and L. D. Schmidt, Science, 1993, 259, 343–346. 32 J. F. Haw, J. B. Nicholas, W. G. Song, F. Deng, Z. K. Wang, T.
2 N. M. Laurendeau, Prog. Energy Combust. Sci., 1978, 4, Xu and C. S. Heneghan, J. Am. Chem. Soc., 2000, 122, 4763–4775.
221–270. 33 W. Song, J. B. Nicholas and J. F. Haw, J. Phys. Chem. B,
3 W. Y. Wen, Catal. Rev.: Sci. Eng., 1980, 22, 1–28. 2001, 105, 4317–4323.
4 M. Asadullah, S. Ito, K. Kunimori, M. Yamada and K. 34 W. Dai, C. Wang, M. Dyballa, G. Wu, N. Guan, L. Li, Z. Xie
Tomishige, J. Catal., 2002, 208, 255–259. and M. Hunger, ACS Catal., 2015, 5, 317–326.
5 D. Sutton, B. Kelleher and J. R. H. Ross, Fuel Process. 35 P. W. Goguen, T. Xu, D. H. Barich, T. W. Skloss, W. Song, Z.
Technol., 2001, 73, 155–173. Wang, J. B. Nicholas and J. F. Haw, J. Am. Chem. Soc.,
6 I. M. Dahl and S. Kolboe, Catal. Lett., 1993, 20, 329–336. 1998, 120, 2650–2651.
7 I. M. Dahl and S. Kolboe, J. Catal., 1994, 149, 458–464. 36 C. Wang, Y. Y. Chu, A. M. Zheng, J. Xu, Q. Wang, P. Gao,
8 I. M. Dahl and S. Kolboe, J. Catal., 1996, 161, 304–309. G. D. Qi, Y. J. Gong and F. Deng, Chem. – Eur. J., 2014, 20, 1–13.
9 S. Kolboe, Acta Chem. Scand., Ser. A, 1986, 40, 711–713. 37 M. J. Wulfers and F. C. Jentoft, ACS Catal., 2014, 4,
10 W. Wang, A. Buchholz, M. Seiler and M. Hunger, J. Am. 3521–3532.
Chem. Soc., 2003, 125, 15260–15267. 38 J. Li, Y. Wei, J. Chen, P. Tian, X. Su, S. Xu, Y. Qi, Q. Wang, Y.
11 D. Lesthaeghe, V. Van Speybroeck, G. B. Marin and M. Zhou and Z. Liu, J. Am. Chem. Soc., 2012, 134, 836–839.
Waroquier, Angew. Chem., Int. Ed., 2006, 45, 1714–1719. 39 S. Xu, A. Zheng, Y. Wei, J. Chen, J. Li, Y. Chu, M. Zhang, Q.
12 S. Wang, Z. Wei, Y. Chen, Z. Qin, M. Dong, W. Fan and J. Wang, F. Deng and Z. Liu, Angew. Chem., Int. Ed., 2013, 52,
Wang, ACS Catal., 2015, 5, 1131–1144. 11564–11568.
13 S. Wang, Y. Chen, Z. Wei, Z. Qin, J. Chen, H. Ma, M. Dong, 40 K. Hemelsoet, J. Van der Mynsbrugge, K. De Wispelaere, M.
J. Li, W. Fan and J. Wang, J. Phys. Chem. A, 2014, 118, Waroquier and V. Van Speybroeck, ChemPhysChem, 2013, 14,
8901–8910. 1526–1545.
14 D. Lesthaeghe, V. Van Speybroeck, G. B. Marin and M. 41 M. Bjørgen, F. Joensen, K. P. Lillerud, U. Olsbye and S.
Waroquier, Chem. Phys. Lett., 2006, 417, 309–315. Svelle, Catal. Today, 2009, 142, 90–97.
15 W. G. Song, J. F. Haw, J. B. Nicholas and C. S. Heneghan, 42 S. Svelle, U. Olsbye, F. Joensen and M. Bjørgen, J. Phys.
J. Am. Chem. Soc., 2000, 122, 10726–10727. Chem. C, 2007, 111, 17981–17984.
16 Ø. Mikkelsen, P. O. Ronning and S. Kolboe, Microporous 43 B. Arstad and S. Kolboe, J. Am. Chem. Soc., 2001, 123,
Mesoporous Mater., 2000, 40, 95–113. 8137–8138.
17 T. Mole and G. Bett, J. Catal., 1983, 84, 435–445. 44 B. Arstad and S. Kolboe, Catal. Lett., 2001, 71, 209–212.
18 T. Mole and J. A. Whiteside, J. Catal., 1983, 82, 261–266. 45 J. Li, Y. Wei, J. Chen, S. Xu, P. Tian, X. Yang, B. Li, J. Wang
19 A. Sassi, M. A. Wildman, H. J. Ahn, P. Prasad, J. B. Nicholas and Z. Liu, ACS Catal., 2015, 5, 661–665.
and J. F. Haw, J. Phys. Chem. B, 2002, 106, 2294–2303. 46 H. Fang, A. Zheng, J. Xu, S. Li, Y. Chu, L. Chen and F. Deng,
20 W. G. Song, J. B. Nicholas, A. Sassi and J. F. Haw, Catal. J. Phys. Chem. C, 2011, 115, 7429–7439.
Lett., 2002, 81, 49–53. 47 Y. Chu, X. Sun, X. Yi, L. Ding, A. Zheng and F. Deng, Catal.
21 R. F. Sullivan, C. J. Egan, G. E. Langlois and R. P. Sieg, J. Am. Sci. Technol., 2015, 5, 3507–3517.
Chem. Soc., 1961, 83, 1156–1160. 48 C.-M. Wang, Y.-D. Wang, Y.-J. Du, G. Yang and Z.-K. Xie,
22 M. Bjørgen, U. Olsbye, D. Petersen and S. Kolboe, J. Catal., Catal. Sci. Technol., 2015, 5, 4354–4364.
2004, 221, 1–10. 49 C. Wang, J. Xu, G. Qi, Y. Gong, W. Wang, P. Gao, Q. Wang,
23 S. Ilias and A. Bhan, J. Catal., 2014, 311, 6–16. N. Feng, X. Liu and F. Deng, J. Catal., 2015, 332, 127–137.
24 R. M. Dessau, J. Catal., 1986, 99, 111–116. 50 C. Wang, X. Yi, J. Xu, G. Qi, P. Gao, W. Wang, Y. Chu, Q.
25 M. Bjørgen, S. Svelle, F. Joensen, J. Nerlov, S. Kolboe, F. Wang, N. Feng, X. Liu, A. Zheng and F. Deng, Chem. – Eur.
Bonino, L. Palumbo, S. Bordiga and U. Olsbye, J. Catal., J., 2015, 21, 12061–12068.
2007, 249, 195–207. 51 C. Wang, Q. Wang, J. Xu, G. Qi, P. Gao, W. Wang, Y. Zou, N.
26 S. Svelle, F. Joensen, J. Nerlov, U. Olsbye, K. P. Lillerud, S. Feng, X. Liu and F. Deng, Angew. Chem., Int. Ed., 2016, 55,
Kolboe and M. Bjørgen, J. Am. Chem. Soc., 2006, 128, 2553–2557.
14770–14771. 52 B. Smit and T. L. M. Maesen, Nature, 2008, 451, 671–678.

This journal is © The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

53 D. Lesthaeghe, B. De Sterck, V. van Speybroeck, G. B. Marin Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li,
and M. Waroquier, Angew. Chem., Int. Ed., 2007, 46, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L.
1311–1314. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J.
54 C.-M. Wang, Y.-D. Wang, Y.-J. Du, G. Yang and Z.-K. Xie, Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H.
Catal. Sci. Technol., DOI: 10.1039/c5cy01419k. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F.
55 J. Li, Z. Wei, Y. Chen, B. Jing, Y. He, M. Dong, X. Li, Z. Qin, Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin,
J. Wang and W. Fan, J. Catal., 2014, 317, 277–283. V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari,
56 J. Liu, Z. F. Liu, G. Feng and D. Kong, J. Phys. Chem. C, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N.
2014, 118, 18496–18504. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V.
57 K. Hemelsoet, A. Nollet, V. Van Speybroeck and M. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E.
Published on 13 April 2016. Downloaded by University of Georgia on 28/04/2016 17:41:44.

Waroquier, Chem. – Eur. J., 2011, 17, 9083–9093. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli,
58 B. Boekfa, P. Pantu, M. Probst and J. Limtrakul, J. Phys. J. W. Ochterski, R. L. Martin, K. Morokuma, V. G.
Chem. C, 2010, 114, 15061–15067. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S.
59 V. Van Speybroeck, J. Van der Mynsbrugge, M. Vandichel, K. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V.
Hemelsoet, D. Lesthaeghe, A. Ghysels, G. B. Marin and M. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision E.01,
Waroquier, J. Am. Chem. Soc., 2011, 133, 888–899. Gaussian, Inc., Wallingford CT, 2009.
60 C. Baerlocher, W. M. Meier and D. H. Olson, Atlas of Zeolite 68 L. Goerigk and S. Grimme, Phys. Chem. Chem. Phys.,
Framework Types, fifthed, Elsevier, Amsterdam, 2001. 2011, 13, 6670–6688.
61 International Zeolite Association: Structural Databases, 69 J. D. Chai and M. Head-Gordon, Phys. Chem. Chem. Phys.,
<http://www.izastructure.org/databases> (accessed 17.01.14). 2008, 10, 6615–6620.
62 B. M. Lok, C. A. Messina, R. Lyle Patton, R. T. Gajek, T. R. 70 J. Van der Mynsbrugge, J. De Ridder, K. Hemelsoet, M.
Cannan and E. M. Flanigen, J. Am. Chem. Soc., 1984, 106, Waroquier and V. Van Speybroeck, Chem. – Eur. J., 2013, 19,
6092–6093. 11568–11576.
63 H. Fujita, T. Kanougi and T. Atoguchi, Appl. Catal., A, 71 D. M. Ruthven and B. K. Kaul, Ind. Eng. Chem. Res.,
2006, 313, 160–166. 1993, 32, 2047–2052.
64 S. R. Lonsinger, A. K. Chakraborty, D. N. Theodorou and 72 S. Svelle, M. Visur, U. Olsbye and M. Bjørgen, Top. Catal.,
A. T. Bell, Catal. Lett., 1991, 11, 209–217. 2011, 54, 897–906.
65 S. Sklenak, J. Dědeček, C. Li, F. Gao, B. Jansang, B. Boekfa, 73 T. R. Forester and R. F. Howe, J. Am. Chem. Soc., 1987, 109,
B. Wichterlová and J. Sauer, Chem. Commun., 2008, 909–920. 5076–5082.
66 P. Feng, X. Bu and G. D. Stucky, Nature, 1997, 388, 735–741. 74 R. Y. Brogaard, R. Henry, Y. Schuurman, A. J. Medford, P. G.
67 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, Moses, P. Beato, S. Svelle, J. K. Norskov and U. Olsbye,
M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. J. Catal., 2014, 314, 159–169.

Catal. Sci. Technol. This journal is © The Royal Society of Chemistry 2016

You might also like