You are on page 1of 18

Journal of Catalysis 330 (2015) 28–45

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Reaction path analysis for 1-butanol dehydration in H-ZSM-5 zeolite:


Ab initio and microkinetic modeling
Mathew John, Konstantinos Alexopoulos, Marie-Françoise Reyniers ⇑, Guy B. Marin
Laboratory for Chemical Technology, Ghent University, Technologiepark 914, B-9052 Gent, Belgium

a r t i c l e i n f o a b s t r a c t

Article history: Dispersion corrected periodic density functional theory (DFT) is used to construct a microkinetic model
Received 27 March 2015 for 1-butanol dehydration in H-ZSM-5. The latter is applied to determine the effect of reaction conditions
Revised 3 June 2015 on dehydration rates, product selectivity and dominant reaction pathway. The consecutive reaction
Accepted 3 July 2015
scheme of 1-butanol dehydration to ether followed by ether decomposition offers lower energy barriers
Available online 17 July 2015
as compared to the direct conversion of 1-butanol to 1-butene. The direct dehydration of 1-butanol to
1-butene occurs via an E2 (anti) elimination at low 1-butanol partial pressure and shifts to a 1-butanol
Keywords:
assisted 1,2-syn-elimination with increasing 1-butanol partial pressure. The ether formation reaction
DFT
Dispersion energy
proceeds via an SN2-type nucleophilic substitution mechanism, involving substitution of the –OH2 group
Bio-butanol of the protonated alcohol by 1-butanol, while ether decomposition predominantly occurs via a
Dehydration 1,2-syn-elimination mechanism. The effect of reaction conditions viz. reaction temperature, site time,
Zeolite 1-butanol and water partial pressure is studied. The reaction conditions govern the coverage of key sur-
Reaction mechanism face species which in turn has a significant role in determining the dominant reaction mechanism and
Microkinetic modeling product selectivity. Under industrially relevant conditions, the presence of water has no significant effect
Reaction path analysis on 1-butanol conversion and product selectivity. A higher reaction temperature, higher site time and
lower 1-butanol partial pressure favor a higher 1-butene selectivity.
Ó 2015 Elsevier Inc. All rights reserved.

1. Introduction accompanied by side reactions such as oligomerization, cracking


and aromatization [18]. Therefore design of a selective catalyst
Being a potential alternative to fossil fuels, biofuels are of prime remains challenging. The first step toward catalyst design is to gain
interest for the development of a sustainable future [1]. First gen- in-depth understanding of the underlying reaction mechanism.
eration bio-fuels such as bio-ethanol are considered as harbingers Earlier efforts to understand the reaction mechanism relied mostly
to the growth in the area of bio-energy. With recent studies indi- upon experimental kinetic and FTIR studies. Based on their kinetic
cating health hazards [2,3] and drawbacks associated with the studies, Olaofe and Yue [9] proposed an E1 mechanism (involving
usage of ethanol as a fuel, much attention is being paid toward Wagner–Meerwein rearrangement of a carbenium ion) for
the usage of higher alcohols such as bio-butanol [4]. Considering 1-butanol dehydration on zeolites. A similar carbenium ion based
the future growth prospects of bio-butanol, it is imperative to look mechanism was used by Zhang et al. [19] to exemplify their reac-
into the various aspects of utilizing it as a feedstock for the chem- tion scheme for conversion of 1-butanol to iso-butene. On the
ical industry. Zeolite catalyzed selective dehydration of other hand, Makarova et al. [17], through their FTIR spectroscopy
bio-butanol to butene could provide an ideal platform for the pro- and their kinetic studies at low pressures (less than 1 kPa) and
duction of such an apt feedstock. Butene produced from butanol temperatures (380–460 K), suggested a reaction mechanism
dehydration could serve as a building block for several specialty involving butoxy intermediate for 1-butanol dehydration reactions
chemicals [5], commodity polymers [6] and fuel additives [7,8]. in ZSM-5 zeolite. They observed the formation of both di-butyl
The catalytic dehydration of butanol has been extensively stud- ether and butene in their dehydration studies. Their kinetic exper-
ied over Brønsted acid sites of several solid acid catalysts, such as iments on ZSM-5 samples of different crystallite size (<0.1–20 lm)
zeolites [9–19], amorphous alumino-silicates [12,17] and polyox- revealed absence of any diffusion limitation for the reactant and
ometalate (POM) clusters [20–24]. The dehydration reaction is the product molecules. Their comparative study for dehydration
of 1-butanol and di-1-butyl ether indicated similar rates for pro-
⇑ Corresponding author. Fax: +32 9 331 1759. duction of butene from both reactants. This could be associated
E-mail address: MarieFrancoise.Reyniers@ugent.be (M.-F. Reyniers). with the possibility of the ether being an intermediate product in

http://dx.doi.org/10.1016/j.jcat.2015.07.005
0021-9517/Ó 2015 Elsevier Inc. All rights reserved.
M. John et al. / Journal of Catalysis 330 (2015) 28–45 29

the path of dehydration of butanol to butene. Overall, there is a


lack of consensus in the literature on the butanol dehydration
mechanism within the zeolite pores.
Theoretical studies could provide significant insight into the
reaction mechanism but there are very few such studies reported
for butanol dehydration in zeolites or over the Brønsted acid sites
of other solid acid catalysts. Macht et al. [20,23,24] studied
2-butanol dehydration on POM clusters via both density functional
theory (DFT) and experimental kinetic studies. Their DFT study
corroborated the formation of a butoxy intermediate, which pro-
duces butene on deprotonation. They found that the reaction fol-
lows an E1 mechanism, with elimination of a water molecule via
a carbenium ion like transition state. However, the dehydration
reaction on the surface of the POM clusters does not involve con-
finement effects, as observed within the zeolite pore structure.
Hence, one cannot disregard the possibility of an alternative reac-
tion mechanism for the latter case. A recent DFT study by
Prestianni et al. [25] indicated a concerted E2 b-elimination mech- Fig. 1. Location of the Brønsted acid site in the unit cell of the H-ZSM-5 structure.
anism for the dehydration of 2-propanol on H-ZSM-5 and H–Y
zeolite.
Owing to the disparity in the available literature regarding the correlation energies were calculated on the basis of the generalized
underlying reaction mechanism, it is imperative to carry out a gradient approximation (GGA) according to Perdew, Burke and
detailed investigation of possible reaction mechanisms. Herein, Ernzerhof (PBE) [37]. Brillouin zone sampling was restricted to
we present a DFT study of possible reaction mechanisms for the C-point. A maximum force convergence criterion of
1-butanol dehydration in H-ZSM-5 zeolite. A comprehensive inves- 0.02 eV Å1 was used and each self-consistency loop was iterated
tigation of the effect of reaction conditions on the reaction rates until a convergence level of 108 eV was achieved. Dispersive cor-
and product selectivity is done using a reaction path analysis. rections for the van der Waals interactions were included by add-
Our study clearly elicits that the dominant reaction mechanism ing a pairwise interaction term to the Kohn–Sham energy using the
is strongly dependent on the reaction conditions. DFT-D2 approach proposed by Grimme [38] and extended by
Kerber et al. [39] for periodic PBE calculations. Although systematic
deviations may be observed due to the overestimation of the dis-
2. Theory
persion interaction [40–43], DFT-D2 has been widely applied for
the theoretical investigation of adsorption [44–47] and reaction
2.1. Catalyst model
in zeolites [47,48] and is known to provide reasonably accurate
results [41,49]. The electronic charge on atoms and fragments
An orthorhombic H-ZSM-5 framework structure (Pnma symme-
was calculated using Bader analysis [50] as implemented by
try, unit cell – HAlSi95O192) was used for the studies. The zeolite
Henkelman et al. [51] Transition state search was performed using
acid site was created by replacing a Si atom with an Al atom and
Nudged Elastic Band (NEB) [52] and dimer [53,54] calculations. The
adding a proton to the adjacent oxygen atom. The MFI orthorhom-
Nudged Elastic Band (NEB) method was used to find an initial
bic unit cell has twelve different tetrahedral sites and the replace-
guess for the minimal energy path (MEP), which was used as a
ment of a Si atom requires selection of an appropriate site. The
starting point for the dimer calculations.
location of the Al atom in the unit cell was reported to be
non-random [26] and to depend on the zeolite synthesis conditions
[27]. Nevertheless, few other studies specify the T12 site to be a
2.2.2. Frequency calculations
preferred location for the Al atom associated with the Brønsted
Normal mode analysis was performed using a Partial Hessian
acid site [28,29]. Moreover, when the Al atom is located at the
Vibrational Analysis (PHVA), relaxing the T5 cluster (HAl(SiO4)4)
T12 site, the Brønsted acid site is accessible to molecules located
of the zeolite framework and the adsorbate molecule for the
in both the straight and sinusoidal channels and could accommo-
numerical Hessian calculation. Previous studies for physisorption
date larger species. Accordingly, the T12 tetrahedral site was
and chemisorption in zeolite have shown that the partial hessian
selected for the replacement of a Si atom with Al.
approach leads to a marginal difference in the result as compared
Likewise, for the positioning of the proton one needs to select
to a Full Hessian Vibration Analysis (FVHA) [55]. Although very
one of the oxygen atoms bound to the Al atom. The study by
stringent optimization (maximum force criterion of 0.02 eV Å1)
Svelle et al. [30] has shown that the proton prefers the oxygen
and electronic convergence (self-consistency loop convergence cri-
atom (Ob) adjacent to the channel intersection rather than the oxy-
terion of 108 eV and energy cutoff of 600 eV) criteria have been
gen atom (Oa) at the channel intersection. Accordingly, the proton
used, spurious imaginary frequencies were still present for very
associated with the Brønsted acidity was placed at the oxygen
few cases, namely C4, TS-1, TS-6 and TS-9 attributed to loosely
atom (Ob) adjacent to the channel intersection as shown in Fig. 1.
bonded species present in these complexes.
The low lying frequencies (<50 cm1) associated with the frus-
2.2. Computational details trated motions of the surface bound species (such as translation or
rotation of the molecule within the zeolite pore structure) could
2.2.1. Electronic energy calculations lead to significant error in the entropy calculations [56–59]. A more
Dispersion corrected periodic DFT calculations were performed accurate estimation of the entropic contributions could be
with the Vienna Ab Initio Simulation Package (VASP) using plane obtained by accounting for anharmonicities by detailed scanning
wave basis sets [31–34]. The electron–ion interactions were of the potential energy surface [60,61], but this would require sig-
described using the projector-augmented wave (PAW) method nificant computational efforts for large systems. Another approach
[35,36] with a plane-wave energy cutoff of 600 eV. The exchange to treat the low lying modes is the use of a frequency cutoff
30 M. John et al. / Journal of Catalysis 330 (2015) 28–45

[44,56,57,62]. De Moor et al. [62] studied the entropy contributions The microkinetic model assumes absence of any diffusion limi-
of these frequencies for alkanes and alkenes in FAU zeolite and tation for reactant and product species. This assumption is consis-
suggested the replacement of these spurious frequencies with tent with the experimental results of Makarova et al. [17], where
50 cm1. Therefore, in order to obtain consistent results, the spuri- they studied the effect of ZSM-5 crystallite size on the reaction
ous imaginary frequencies and low-lying frequencies were rates for dehydration of 1-butanol and di-1-butyl ether. The above
replaced by normal modes of 50 cm1 [62]. mentioned set of ordinary differential equations is integrated using
the LSODA module of ODEPACK [64]. The investigated reaction
2.2.3. Statistical thermodynamics conditions are partial pressure of 1-butanol (0.01–100 kPa) and
Standard enthalpies, entropies, Gibbs free energies, adsorption water (0–40 kPa), site time (NH+/FBuOH,0 = 0–200 mol H+ s/
and reaction equilibrium coefficients (K) are obtained from total mol BuOH0) and reaction temperature (400–460 K).
partition functions by statistical thermodynamics calculations
[63]. 3. Results and discussion
Reaction equilibrium coefficients, K, for elementary reactions
are calculated as follows: 3.1. Adsorption of 1-butanol in H-ZSM-5
Q  
Q ðN; V; TÞ DEr
KðTÞ ¼ Qi i exp  ð1Þ The adsorption of the 1-butanol molecule on the Brønsted acid
j Q j ðN; V; TÞ RT
site at the intersection of the channels is studied first. The butanol
where i and j denote products and reactants respectively. DEr is the molecule can be adsorbed either in the straight or in the sinusoidal
change in electronic energy at 0 K (including the zero-point vibra- channel of the H-ZSM-5 framework. Previous theoretical studies on
tional energy) of the reaction and Q the total partition function. physisorption and protonation of alcohols in H-ZSM-5 have shown
The electronic energy from the DFT calculation along with the fre- that there is no energetic preference for adsorption of 1-butanol in
quencies obtained from the vibrational analysis is used for the sta- the straight or sinusoidal channel of the zeolite [45,46]. Herein, for
tistical thermodynamic calculation (see supporting information). the adsorption of a single molecule of 1-butanol, the molecule is
The partition functions for the gas-phase species included vibra- considered to be situated in the straight channel, while the second
tional, rotational and translational degrees of freedom, while only molecule is adsorbed in the sinusoidal channel. Representative
the vibrational contributions were taken into account for the sur- physisorbed and protonated states for 1-butanol in H-ZSM-5 zeo-
face species. lite are depicted in Fig. 2. The geometrical parameters and Bader
The reaction rate coefficients of elementary reaction steps are charges (of the butanol molecule and the protonated butanol frag-
calculated on the basis of transition state theory: ment) for the different adsorbed forms of 1-butanol are summa-
! rized in supporting information Table S1.
kB T Q TS ðN; V; TÞ Ez
kTST ðTÞ ¼ exp  ð2Þ
h Q R ðN; V; TÞ RT 3.1.1. Physisorbed 1-butanol
During physisorption, the oxygen atom of the butanol molecule
where kB is Boltzmann constant, h is Planck constant and Eà is the readily forms a strong hydrogen bond with the Brønsted acid site
electronic activation barrier at 0 K (including the zero-point vibra- of the zeolite. As illustrated in Fig. 2, depending on the orientation
tional energy). QTS and QR denote the total partition functions of of the adsorbed molecule, it can form a second hydrogen bond with
the transition and reactant state respectively. Arrhenius the oxygen (Oc) bridging two silica atoms, forming P1 or with the
pre-exponential factors (A) and activation energies (Ea) are obtained oxygen atom adjacent to the Al atom (Oa), forming P2. The stan-
by regression of Eq. (2) in the temperature range of 300–800 K. dard adsorption enthalpy/entropy and adsorption equilibrium
The adsorption occurs without any activation barrier. Hence, coefficients are reported in Table 1. The standard adsorption
the reaction rate coefficient for the adsorption step is calculated enthalpy for P2 is marginally higher than that of P1. Both P1 and
as kads = kBT/h, while the reaction rate coefficient for the desorption P2 have similar values for adsorption entropy. According to the
step is calculated from thermodynamic consistency as Bader analysis, the butanol fragment of P1 has a total charge of
kdes = kads/Kads. +0.108 as compared to +0.117 for P2. This can be rationalized on
the basis of a strong hydrogen bond (bond length 133.8 pm)
2.3. Microkinetic model between the oxygen atom of butanol (O1) and the zeolite proton
(Hb) in P2, giving it partial characteristics of the protonated species
A microkinetic modeling approach has been used to carry out as seen in the next section. On the other hand, a relatively weaker
reaction path analysis and to study the effect of reaction conditions interaction of the alcohol with the zeolite Brønsted acid site leads
(temperature, site time and pressure). The net production fre- to a lower partial charge on butanol in the P1 monomer. These
quency, of any considered component i is obtained by linear sum- results for physisorption are also in line with the previous studies
mation of the rates of the elementary steps in which the by Nguyen et al. [45,46].
component i is involved:
X
TOF i ¼ mij rj ð3Þ 3.1.2. Protonated 1-butanol
j The protonation of the butanol molecule on the Brønsted acid
site of the zeolite is a key step in the butanol dehydration reaction.
in which mij is the stoichiometric coefficient of component i in the The monomeric protonated form, M1, is hydrogen bonded to two
elementary step j. An isothermal plug flow reactor model is used zeolite oxygen atoms which are adjacent to the Al atom. This is a
for the reactor simulations: highly exothermic adsorption with a standard adsorption enthalpy
(DHoads) of 146 kJ/mol. The adsorption equilibrium coefficient (Kads)
dF i
¼ C t TOF i ð4Þ at 400 K for M1 is an order of magnitude higher than that for the
dW
physisorbed forms (P1 and P2). However, as seen from the support-
where Fi is the molar flow rate of component i (mol/s), W the cata- ing information (Fig. S1) the difference in stability of physisorbed
lyst mass (kg), Ct the acid site concentration (mol H+/kg), TOFi the and protonated monomer is somewhat more pronounced when
net production frequency of component i (moli/mol H+/s). using DFT-D2 as compared to vdW-DF2. The influences of electronic
M. John et al. / Journal of Catalysis 330 (2015) 28–45 31

Fig. 2. Adsorbed forms of 1-butanol in H-ZSM-5. (a) P1 – physisorbed butanol configuration 1, (b) P2 – physisorbed butanol configuration 2, (c) M1 – protonated monomer,
(d) D1 – protonated dimer, (e) D2 – protonated monomer next to physisorbed butanol. Color code: silicon – light blue, aluminum – pink, oxygen – red, hydrogen – white,
carbon – gray, hydrogen bonds (distance < 250 pm) – light green. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of
this article.)

Table 1 molecule is a more suitable candidate for the elimination and sub-
Thermodynamics for the adsorbed monomer and dimer forms of 1-butanol in H-ZSM- stitution reactions that follow.
5 zeolite. DEDFT – electronic adsorption energy at 0 K (kJ/mol), DED – dispersive
contribution to adsorption energy (kJ/mol), DEDFT-D – DFT-D adsorption energy at 0 K
(kJ/mol), standard adsorption enthalpy (DHoads in kJ/mol), adsorption entropy (DSoads in
3.1.3. Adsorbed butanol dimer
J/mol/K), and equilibrium coefficient at 400 K (K400K
ads in 10
2
kPa1 and 104 kPa2 for
the adsorption of one and two molecules of butanol, respectively). If a second 1-butanol molecule is adsorbed at the sinusoidal
channel, there is a complete abstraction of the zeolite proton,
DEDFT DED DEDFT-D DHoads DSoads K400K
ads
which is shared by the two butanol molecules. This adsorption is
P1 70 65 135 130 178 5.3  107 also highly exothermic (DHoads = 272 kJ/mol) and leads to the for-
P2 65 70 135 132 178 8.4  107
mation of the dimer D1, stabilized by strong hydrogen bonds
M1 77 72 149 146 192 1.1  109
D1 110 173 283 272 375 8.0  1015
between the two molecules and with the zeolite surface. The strong
D2 87 154 241 229 378 1.4  1010 binding with the zeolite surface leads to a significant entropy loss
(DSoads = 375 J/mol/K). The formation of such stable dimer interme-
diates on Brønsted acid sites has been reported by Macht et al.
[20,23,24]. As illustrated in Fig. 2, the dimer D1 can undergo a
re-arrangement via cleavage of the hydrogen bond between the pro-
and dispersive forces along with enthalpy and entropy contributions
tonated alcohol and the butanol molecule to form a thermodynam-
for the adsorption of the butanol molecule are tabulated in Table 1. A
ically less favorable configuration, D2 (see Table 1). A significant
significant contribution of dispersive forces in the stabilization of the
fraction of the overall adsorption energy comes from the dispersive
adsorbed butoxonium is observed. The Bader charge on the proto-
contribution, since the bulky dimer molecules (D1 and D2) feel a
nated alcohol is found to be +0.769, with an equivalent negative
large dispersive stabilization. A comparison can be made based on
charge on the deprotonated zeolite framework. A stabilization effect
the energetics of the different species reported in Table 1. The dimer
by this ion pair interaction is thus observed in this study. These
D1 has an adsorption equilibrium coefficient which is a few orders of
results are consistent with previous studies [45,46].
magnitude higher than all the other adsorbed reactant species
Our results show that there is an elongation of the Ca–O1 bond
(KadsD1  KadsD2  KadsM1 > KadsP2  KadsP1).
distance from 143.8 to 148.8 pm. The protonation of the alcohol
molecule weakens the strength of the C–O bond and facilitates
the heterolytic cleavage of the C–O bond. Similar results have been 3.2. Reaction pathways for 1-butanol dehydration in H-ZSM-5
reported for alkyl halides (R–X), where the protonation was found
to favor the heterolytic scission of the C–X bond [65]. As inferred The adsorbed 1-butanol molecule can undergo a direct dehy-
by the elongation of the C–O bond, the protonated alcohol dration reaction producing 1-butene (path A) or react in a
32 M. John et al. / Journal of Catalysis 330 (2015) 28–45

the supporting information. A brief definition of the investigated


elimination and substitution mechanisms is available in support-
ing information. Any further reference to a specific reaction step
of the network is done as per the numbering used in Fig. 4. The
standard reaction enthalpy (DHor ), reaction entropy (DSor ),
Arrhenius activation energies and pre-exponential factors, forward
reaction rate coefficients and equilibrium coefficients for each ele-
mentary step are tabulated in Table 3, while the corresponding
reaction energies and activation barriers at 0 K are reported in
Table S4 of the supporting information.

3.2.1. Path A: dehydration of 1-butanol to 1-butene


Five different reaction mechanisms are considered for the direct
conversion of 1-butanol to 1-butene in H-ZSM-5, of which four are
monomolecular and one is bi-molecular. The reaction mechanisms
Fig. 3. Reaction scheme for dehydration of 1-butanol to 1-butene (path A),
dehydration of 1-butanol to di-1-butyl ether (path B) and ether decomposition are classified based on the geometric and electronic features of the
(path C). transition state intermediate. The transition state structures for the
different reaction mechanisms associated with the direct dehydra-
tion of 1-butanol to 1-butene are shown in Fig. 5.
sequential manner to yield di-1-butyl ether (path B) which can fur-
ther decompose to 1-butene and 1-butanol (path C) as shown in 3.2.1.1. Reaction mechanism 1 (E1-like). This mechanism consists of
Fig. 3. the sequence of elementary steps 1, 2 and 3 as illustrated in Fig. 4.
An overview of the reaction network consisting of all the ele- In step 2, the crucial step of the mechanism, the adsorbed mono-
mentary steps considered for the dehydration of butanol in mer M1 undergoes a heterolytic cleavage of the C–O bond forming
H-ZSM-5 zeolite is depicted in Fig. 4. This reaction network con- a 4-ring (O1–Ca–Cb–H4) transition state (TS-1), involving a carbe-
sists of 10 different mechanisms that are involved in the aforemen- nium ion like [C4H9] fragment and a water molecule (see TS-1,
tioned three reaction pathways as shown in Table 2. A detailed Fig. 5). Both hydrogen atoms of the water molecule are hydrogen
electron flow diagram for each mechanism is shown in Fig. S2 of bonded to the zeolite oxygen atoms (Oa and Ob) adjacent to the

Fig. 4. Detailed reaction network for dehydration of butanol in H-ZSM-5.


M. John et al. / Journal of Catalysis 330 (2015) 28–45 33

Table 2
Elementary steps and reaction mechanisms for the butanol dehydration reaction. Non-equilibrated steps are indicated with red stoichiometric numbers.

Table 3
Standard reaction enthalpy (kJ/mol), reaction entropy (J/mol/K), activation energy (kJ/mol), pre-exponential factor (s1), forward reaction rate coefficient kf (s1) at 400 K and
equilibrium coefficient at 400 K (102 kPa1, 102 kPa or dimensionless for adsorption, desorption and surface transformation, respectively) for the elementary steps (numbered as
indicated in Fig. 4).

Elementary steps DHor DSor Ea(f) Af kf(400K) Keq(400K)


(R1) 1-BuOH(g) + ⁄ M M1 146 192 – – – 1.1  109
(R2) M1 M W + 1-Butene(g) 107 200 176 1.1  1015 1.1  108 2.7  104
(R3) W M H2O(g) + ⁄ 85 150 – – – 5.0  104
(R4) M1 M C1 74 78 139 2.9  1014 1.9  104 2.9  106
(R5) C1 M W + 1-Butene(g) 34 122 – – – 9.3  101
(R6) M1 M M2 82 5 – – – 1.1  1011
14
(R7) M2 M 1-Butene⁄ + H2O(g) 28 199 53 9.0  10 1.2  108 5.7  106
(R8) 1-Butene⁄ M 1-Butene(g) + ⁄ 83 155 – – – 2.2  103
(R9) M2 M Butoxy + H2O(g) 22 165 49 3.4  1014 1.3  108 5.1  105
(R10) Butoxy M 1-Butene⁄ 6 35 93 3.7  1013 2.3  101 1.1  101
(R11) M1 + BuOH(g) M D1 126 183 – – – 7.4  106
(R12) D1 M D2 43 3 – – – 1.7  106
14 2
(R13) D2 M C2 + 1-Butene(g) 69 165 118 2.8  10 9.8  10 4.5  101
(R14) C2 M M1 + H2O(g) 61 179 – – – 2.5  101
(R15) D2 M DBE⁄ + H2O(g) 15 157 92 1.4  1014 1.4  102 1.5  106
(R16) DBE⁄ M DBE(g) + ⁄ 191 209 – – – 1.1  1014
(R17) Butoxy + BuOH(g) M C3 95 173 – – – 2.0  103
(R18) C3 M DBE⁄ (Sn2) 77 16 61 3.1  1012 3.4  104 1.8  109
(R19) C3 M DBE⁄ (Sn1) 77 16 111 9.2  1013 2.8  102 1.8  109
(R20) DBE⁄ M C4 102 51 140 2.5  1014 1.3  104 2.2  1011
(R21) C4 M 1-Butene⁄ + BuOH(g) 76 173 – – – 1.4  101
(R22) DBE⁄ M DBE2 63 9 – – – 1.9  108
(R23) DBE2 M 1-Butene⁄ + BuOH(g) 115 215 84 1.6  1013 3.0  106 3.1  1012

Al atom. Meanwhile the atom (O1) of the adsorbed water molecule bond is completely broken while the Cb–H4 still remains intact
attacks the b-hydrogen (H4) of the C4H9 fragment leading to the indicating an E1-like mechanism. Although a carbenium ion like
formation of a physisorbed 1-butene and an adsorbed hydronium TS is observed, formation of a stable carbocation intermediate
ion (W). Thereafter, the butene molecule desorbs leaving the characteristic of a pure E1 mechanism is not observed. This can
adsorbed hydronium ion. The hydronium undergoes deprotonation be attributed to the high reactivity of the primary carbocation.
and the desorption of a water molecule regenerates the zeolite Overall, a late transition state is evident from the fact that the geo-
Brønsted acid site. The step leading to the formation of butene metric configuration (e.g. extent of Ca–O1 bond breakage and other
and hydronium from M1 is an activated process with an activation bond distances/angles) of the transition state is much closer to the
energy of 176 kJ/mol. product (see supporting information, Table S2).
In the transition state (TS-1, Fig. 5) geometry, Ca–O1 and Cb–H4 Further insight into the nature of the transition state was
bond lengths are 231 and 118 pm, respectively. Herein, the C–O obtained via a Bader charge analysis. The Bader charge for the
34 M. John et al. / Journal of Catalysis 330 (2015) 28–45

Fig. 5. Transition state structures (views along the sinusoidal channel) for 1-butanol dehydration to 1-butene (path A): (1) TS-1, 4-ring E1-like transition state, (2) TS-2, syn-
elimination, (3) TS-3, E-2 (anti) elimination, (4) TS-4, SN2 substitution, (5) TS-5, deprotonation/decomposition of 1-butoxide and (6) TS-6, butanol assisted syn elimination.
Color code: silicon – light blue, aluminum – pink, oxygen – red, hydrogen – white, carbon – gray, hydrogen bonds – green dashed lines, bond breaking/forming – black dashed
lines. Important interatomic distances are shown in pm. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)

[C4H9] fragment and the water molecule was found to be +0.797 cleavage. As described above, the protonation of butanol to form
and +0.0715 respectively, while the zeolite framework had an the adsorbed monomer M1 leads to an elongation of the C–O bond
overall charge of 0.8685. The charge on the [C4H9] fragment is making it easier to break the C–O bond as compared to the C–H
comparable with the one reported for the cationic fragment bond and this gives the mechanism an E1-like character.
(+0.85) in a DFT study on E1 elimination of 2-butanol in POM [23]. The total Bader charge on the [C4H9] fragment is +0.6671, but
the major fraction of this charge is associated with the
3.2.1.2. Reaction mechanism 2 (syn-elimination). This dehydration b-hydrogen (+0.3905) which has nearly broken its bond with the
mechanism consists of steps 1, 4, 5 and 3 as illustrated in Fig. 4. b-carbon (Cb). Moreover, the formal charge on the [C4H9] fragment
The formation of adsorbed monomer M1 (step 1) is followed by is much lower than the literature reported values for an E1-type
an activated step (step 4), involving a concerted mechanism, mechanism [23]. Thus, the results of the Bader analysis confirm
wherein there is a simultaneous cleavage of the Ca–O1 bond along that although the TS has some E1 character, it essentially repre-
with an abstraction of the b-hydrogen by the zeolite basic oxygen sents a one step 1,2-syn-elimination reaction.
(Oa). For this concerted reaction to occur, the M1 monomer has to
reorient itself by breaking the H0–Oa hydrogen bond and aligning 3.2.1.3. Reaction mechanism 3 (E2 elimination). This dehydration
the b-hydrogen toward Oa (NEB shown in the supporting informa- mechanism consists of steps 1, 6, 7 and 8. It occurs via an
tion, Fig. S4). The simultaneous cleavage of Ca–O1 and Cb–H4 (step E2-type elimination that requires the atoms or groups involved
4) has activation energy of 139 kJ/mol. This step is followed by a in the reaction to be in the same plane, with antiperiplanar orien-
subsequent desorption of 1-butene (step 5) and water (step 3) tation of the leaving group and the b-hydrogen. To achieve this
respectively. configuration, the protonated monomer (M1) undergoes a
In the transition state (TS-2, Fig. 5) structure, the b-hydrogen re-orientation (step 6), breaking two strong hydrogen bonds and
(H4) and the leaving group (water) have a near syn-coplanar con- forming a new weaker hydrogen bond with the oxygen atom
figuration with a O1–Ca–Cb–H4 dihedral angle of 17.3°. The (Od) bridged between two Si atoms, to form a loosely adsorbed
a-carbon and b-carbon have a near planar configuration with the monomer M2, having the required antiperiplanar configuration.
a-carbon being closer to planarity than the b-carbon (see support- The M2 monomer has a higher positive formal charge of +0.863
ing information, Table S2). The extent of Ca–O1 bond breakage is as compared to +0.769 for the protonated M1 monomer, indicating
found to be more prominent as compared to that of the Cb–H4 that the zeolite proton is more strongly associated with the M2
M. John et al. / Journal of Catalysis 330 (2015) 28–45 35

monomer. On rearrangement to M2, the C–O bond distance is elon- and Oa and assumes a penta-coordinated trigonal bi-pyramidal
gated to 154 pm as compared to 149 pm for M1 facilitating the state, characteristic of a SN2-type substitution reaction.
removal of the –OH2 species. M2 is less stable as compared to In the subsequent step (step 10), the 1-butoxide undergoes a
M1 with a free energy difference of 80 kJ/mol. Thus, M2 repre- de-protonation reaction, in which the zeolite oxygen atom (Ob)
sents a high energy metastable intermediate that fulfills the stere- attacks the b-hydrogen (H4), while the primary carbon atom breaks
ochemistry required for the following elimination reaction. its bond with the zeolite oxygen (Oa). There is a simultaneous elon-
The M2 monomer undergoes an elimination through a con- gation and cleavage of the Ca–Oa and Cb–H4 bonds. In the transition
certed mechanism (step 7), involving simultaneous cleavage of state geometry (TS-5, Fig. 4), the a-carbon and b-carbon have a
the C–O bond and abstraction of the b-hydrogen by the zeolite near-planar configuration with the a-carbon being closer to pla-
basic oxygen (Oa). This step is an activated process and has an acti- narity than the b-carbon. The Ca–Oa and Cb–H4 bond lengths in
vation energy of 53 and 135 kJ/mol, considering M2 and M1 as ref- the transition state are 224 and 119 pm respectively. Again, a late
erence states, respectively. Due to the hydrogen abstraction, the transition state is evident, as the TS structure is close to the pro-
zeolite regains its proton and the loosely bound water molecule duct rather than the reactant (see supporting information,
desorbs leaving 1-butene physisorbed over the Brønsted acid site. Table S2).
Desorption of butene (step 8) regenerates the Brønsted acid site. The Bader charge analysis for TS-4 and TS-5 shows a cationic
In the transition state (TS-3, Fig. 5) geometry, the b-hydrogen nature of the transition state fragments. The Bader charge analysis
(H4) points toward the bridge oxygen atom (Oa) and the leaving allocated a total charge of +0.7581 on the [C4H9–H2O] fragment of
–OH2 group is positioned opposite to it, making a dihedral angle the TS-4 transition state, with the [C4H9] fragment and [H2O] frag-
(O1–Ca–Cb–H4) of 180.9°. Looking at the bond angles, it is evident ment having Bader charges of +0.6353 and +0.1228 respectively.
that both the a-carbon (Ca) and b-carbon (Cb) have a near planar The charge on the [C4H9] fragment of TS-5 is +0.6916, but the
configuration with the a-carbon being closer to planarity than major fraction of this charge is associated with the b-hydrogen
the b-carbon (see supporting information, Table S2). The Ca–O1 (+0.3222). The large positive charge on the b-hydrogen atom (H4)
and Cb–H4 bond lengths in the transition state are 223 and is associated with the elongation of the Cb–H4 bond and the forma-
130 pm respectively. The extent of C–O bond breakage is more pro- tion of a new bond of b-hydrogen (H4) with the zeolitic oxygen
nounced than that of the b-carbon (Cb) and b-hydrogen (H4). The (Ob).
E1 character in the transition state could be explained by the sig-
nificant elongation of the C–O bond in the reactant M2, making it 3.2.1.5. Reaction mechanism 5 (butanol-assisted syn-elimination). In
easier to break the C–O bond as compared to the C–H bond during this mechanism, the adsorption of another butanol molecule on
the reaction step. TS-3 has a Ca–Cb bond distance of 138 pm, which the protonated M1 (step 11) leads to formation of an adsorbed
is similar to that of the physisorbed 1-butene and indicates a late butanol dimer D1, which undergoes a reorientation (step 12) to
transition state. Thus, we can conclude that this mechanism fol- form D2. This is followed by an activated elimination reaction with
lows a one step 1,2-anti-elimination (E2) reaction with E1 desorption of 1-butene (step 13) and a subsequent water desorp-
character. tion (step 14). In step 13, the protonated alcohol of the reoriented
Comparing anti- and syn-elimination (TS-3 vs TS-2), the adsorbed butanol dimer (D2) undergoes a concerted elimination
b-carbon (Cb) is closer to planarity in the anti-elimination. reaction, similar to the monomolecular 1,2-syn-elimination reac-
The anti-elimination has an activation energy of 135 kJ/mol tion (reaction mechanism 2). Herein, the b-hydrogen abstraction
(with respect to M1) which is slightly lower than that for is done by the oxygen atom (O2) of the physisorbed butanol mole-
syn-elimination. cule instead of the basic zeolitic oxygen. This provides a bimolecu-
lar route for the production of butene without involving ether as an
intermediate product.
3.2.1.4. Reaction mechanism 4 (1-butanol to 1-butoxide via SN2 type In the transition state (TS-6, Fig. 5) structure, the b-hydrogen
reaction, followed by deprotonation to butene). In this mechanism, (H4) and the leaving group (water) have a syn-coplanar configura-
the protonated alcohol undergoes a nucleophilic substitution with tion with O1–Ca–Cb–H4 dihedral angle of 0.4°. The a-carbon and
the zeolite oxygen atom (Oa) to form 1-butoxide, which then under- b-carbon both have a near planar configuration with a-carbon
goes deprotonation to give 1-butene. The mechanism involves a being slightly more planar than the b-carbon (see supporting infor-
sequence of elementary steps 1, 6, 9, 10 and 8, of which two steps mation, Table S3). Here again, the extent of C–O bond (bond dis-
(9 and 10) are activated. The SN2-type substitution reaction has tance 218 pm) breakage is found to be more prominent as
also specific stereochemical requirements as was the case for the compared to that of the b-carbon and hydrogen cleavage
E2 anti-elimination reaction. Moreover, it requires simultaneous (125.6 pm).
bond breaking (i.e. C–OH2 bond of the protonated alcohol) and bond The physisorbed butanol and water molecule have a minor
forming (between Ca and Oa of zeolite). The formation of the new charge of +0.1179 and +0.0832 respectively, while the zeolite
bond eases the breaking of the existing bond. To minimize the framework has a charge of 0.8589. The charge on the [C4H9] frag-
repulsion of the electronic clouds of the approaching and departing ment is +0.6578, but a major fraction of this charge is associated
species, the nucleophile (Oa) and the leaving group (–OH2) should with the b-hydrogen (+0.3733) which has nearly broken its bond
align themselves in a direction opposite to each other (dihedral with the b-carbon (Cb). Based on the geometrical aspects and
angle close to 180°) forming a trigonal bi-pyramidal structure. results of the Bader charge analysis, one can conclude that the
The M2 monomer described earlier meets these stereochemical key step of this reaction mechanism follows a one step
requirements associated with the SN2 back-side attack. 1,2-syn-elimination with E1 character.
In this mechanism, M1 rearranges to M2 (step 6), which then
undergoes a nucleophilic substitution reaction to form a surface 3.2.1.6. Comparison of the reaction mechanisms for the direct
bound 1-butoxide (step 9). Herein, the basic oxygen (Oa) of the dehydration of 1-butanol to 1-butene. The Gibbs free energy profile
zeolite surface acts as a nucleophile and attacks the primary carbon for the five different reaction mechanisms for the direct dehydra-
(Ca). The primary carbon atom Ca breaks its bond with the leaving tion of 1-butanol to 1-butene (path A) at 400 K is shown in
group (–OH2) and concurrently forms a new Ca–Oa bond with the Fig. 6. The reaction mechanism having a 4-ring transition state
zeolite surface. A detailed analysis of the transition state structure (TS-1) with E1 character was found to have the highest free energy
(TS-4, Fig. 5) reveals that the Ca carbon atom is equidistant from O1 barrier. The 1,2-anti-elimination (TS-3) is seen to be favored over
36 M. John et al. / Journal of Catalysis 330 (2015) 28–45

Fig. 6. Gibbs free energy profile for different plausible mechanisms of the direct dehydration of 1-butanol to 1-butene in H-ZSM-5 at 400 K.

the 1,2-syn-elimination (TS-2) reaction. The reaction mechanism reaction leading to the formation of protonated ether (step 15) and
involving the butoxide intermediate has two transition states, the desorption of ether (step 16). In step 15, the protonated alcohol
namely TS-4 and TS-5. Here, the reaction of 1-butoxide to part of D2 undergoes a nucleophilic substitution of the –OH2 group
1-butene via TS-5 has a slightly higher free energy barrier as com- with the physisorbed butanol molecule to form a protonated
pared to the reaction of the protonated butanol monomer (M2) to di-butyl ether and a water molecule. During this etherification
1-butoxide via TS-4. With an increase in the temperature, the reaction, the protonated alcohol breaks one of its H-bond (Oa–H0)
entropic contributions make significant changes to the Gibbs free and reorients itself to place the a-carbon (Ca) between the oxygen
energy profile and the free energy barrier at 500 K becomes higher atoms of the leaving group (–OH2) and the physisorbed butanol
for TS-4 as compared to TS-5 (see Fig. S5 of supporting informa- molecule. This is an activated step and the transition state struc-
tion). Among the monomolecular mechanisms for path A (i.e. ture is shown in Fig. 7 (TS-7).
mechanisms 1–4), the one step 1,2-anti-elimination via TS-3 pro- The transition state (TS-7) has a trigonal bipyramidal configura-
vides the lowest free energy barriers with an apparent DGà of tion with the Ca carbon in a penta-coordinated state, which is char-
51 kJ/mol at 400 K (with respect to the gas phase butanol and acteristic of a SN2-type substitution reaction. The b-carbon (Cb)
the zeolite). The conversion of the butanol dimer D2 to C2 remains in a tetrahedral state. The bond distance between Ca–O1
(co-adsorbed butanol and water) and gas-phase 1-butene provides and Ca–O2 is 197 and 218 pm respectively with Ca placed near
a bimolecular mechanism for the direct dehydration of 1-butanol the center of O1 and O2. The O1–Ca–O2 angle for the transition state
to 1-butene. This butanol-assisted syn-elimination reaction offered structure is 162.7°. This orientation of O1–Ca–O2 in the transition
the lowest free energy barrier (apparent DGà of 30 kJ/mol at 400 K) state largely depends on the stabilization of the leaving water
as compared to all other path A mechanisms. It is pertinent to note molecule by hydrogen bonding with the zeolite oxygen.
that although the butanol-assisted syn-elimination offers a lower The Bader charge on [H2O–C4H9–C4H9OH] is +0.8786, with the
free energy barrier at 400 K, it becomes less favorable with an [C4H9] fragment, the [H2O] molecule and the physisorbed alcohol
increase in temperature (see Fig. S5 of supporting information). having +0.582, +0.1875 and +0.1091 respectively. In case of a
For comparison with the experimental results of Makarova et al. SN1 kind of mechanism, we would have expected the net charge
[17], the activation energy for the activated steps was calculated on the [C4H9] fragment to be comparable to its value for an E1
with respect to the adsorbed state M1, since this is expected to mechanism (i.e. +0.797, see mechanism 1). This further corrobo-
be the most abundant reaction intermediate at the low pressure rates that the reaction step is following an SN2 mechanism as
conditions used by Makarova et al. [17]. Relative to M1, the inferred from the geometric conformation of the transition state.
anti-elimination (via TS-3) has an activation energy (Ea,(TS3-M1)) of
135 kJ/mol, which is in rather good agreement with the experi-
3.2.2.2. Reaction mechanism 7 (formation of di-1-butyl ether from 1-
mentally reported value of 138 ± 8 kJ/mol [17].
butoxide and 1-butanol via SN2 type reaction). 1-Butoxide produced
as an intermediate in mechanism 4 (steps 1–6–9, via TS-4) can
3.2.2. Path B: Etherification reaction (dehydration of 1-butanol to di-1- react with another butanol molecule leading to the formation of
butyl ether) protonated ether (DBE⁄) and water. This reaction mechanism
Makarova et al. [17] observed significant formation of involves a sequence of elementary steps, namely the physisorption
di-1-butyl ether in their 1-butanol dehydration studies in of 1-butanol (step 17) alongside the surface bound butoxide spe-
H-ZSM-5. In view of this, three possible mechanisms for the con- cies, the nucleophilic substitution reaction leading to the formation
version of 1-butanol to di-1-butyl ether have been envisaged. of protonated ether (step 18) and the deprotonation and desorp-
One mechanism proceeds via the dehydration of the butanol dimer tion of di-1-butyl ether (step 16). In step 18, the oxygen atom
while the other two mechanisms involve the reaction of physi- (O2) of butanol attacks the a-carbon atom (Ca) of the butoxide,
sorbed butanol with surface bound butoxide species. The transition which is accompanied by the concurrent cleavage of the Ca–Oa
state structures for the different reaction mechanisms associated bond of the butoxide forming protonated ether adsorbed on the
with the dehydration of 1-butanol to di-1-butyl ether are shown zeolite. This is an activated step and the transition state (TS-8) geo-
in Fig. 7. metrical parameters are tabulated in Table S3 of the supporting
information.
3.2.2.1. Reaction mechanism 6 (butanol dimer to ether and water via In the transition state structure (TS-8, Fig. 7), it is seen that the
SN2 type reaction). This reaction mechanism consists of a sequence bond between Ca–Oa is elongated and Ca is placed near the center
of steps, namely, formation of adsorbed 1-butanol dimer D1 (step of O2 and Oa. The bond distance between Ca–O2 and Ca–Oa is 214.3
11), reorientation of D1 to D2 (step 12), a nucleophilic substitution and 208.5 pm respectively. The a-carbon (Ca) acquires a trigonal
M. John et al. / Journal of Catalysis 330 (2015) 28–45 37

Fig. 7. Transition state structures (views at the channel intersection) for the etherification reaction (path B): (1) TS-7, SN2 substitution reaction (mechanism 6, 1-butanol
dimer (D2) to di-1-butyl ether), (2) TS-8, SN2 substitution reaction (mechanism 7, 1-butoxide and 1-butanol to dibutyl ether), (3) TS-9, SN1 like substitution reaction
(mechanism 8, 1-butoxide and 1-butanol to di-1-butyl ether). Color code: silicon – light blue, aluminum – pink, oxygen – red, hydrogen – white, carbon – gray, hydrogen
bonds (distance < 250 pm) – green dashed lines, bond breaking/forming – black dashed lines. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

bipyramidal structure which is characteristic of a SN2-type substi- analysis indicate a higher positive charge of +0.7096 on the
tution reaction while the b-carbon (Cb) remains in a tetrahedral [C4H9] fragment as compared to +0.5937 for the SN2 mechanism
state. The O2–Ca–Oa angle in the transition state configuration is 7. The presence of a carbenium-like fragment, having a pronounced
160°. The Bader charge on the [C4H9–OH–C4H9] fragment is positive charge, is indicative of a SN1-like mechanism.
+0.7482, with the [C4H9] fragment and the physisorbed alcohol
having +0.5937 and +0.1545 respectively. The charge on the 3.2.2.4. Comparison of the reaction mechanisms for the etherification
[C4H9] fragment is consistent with that of the SN2 reaction (mech- reaction of 1-butanol. The Gibbs free energy profile for the three
anism 6, step 15) for the conversion of D2 to protonated ether. different mechanisms of the dehydration of 1-butanol to dibutyl
ether (path B) is shown in Fig. 8. Among the three plausible mech-
3.2.2.3. Reaction mechanism 8 (formation of di-1-butyl ether from anisms, the conversion of the dimer D2 to protonated di-1-butyl
1-butoxide and 1-butanol via SN1 type reaction). This reaction ether and water via a SN2 reaction (TS-7) offered the lowest free
mechanism follows a sequence of elementary steps similar to that energy barrier with an apparent DGà of 5 kJ/mol at 400 K with
of mechanism 7, except that the backside SN2 nucleophilic substi-
tution (step 18) is replaced by a SN1 like substitution reaction (step
19). Herein, the reaction proceeds via the heterolytic cleavage of
the Ca–Oa bond leading to the formation of a carbenium ion like
C4H9 fragment which undergoes a nucleophilic attack by the oxy-
gen atom (O2) of the butanol. However, as seen from Table 3, the
activated step 19 has a significantly higher activation energy com-
pared to step 18 of the SN2 mechanism 7.
In the transition state structure (TS-9, Fig. 7), it is seen that the
bond between Ca–O2 is not formed yet with an interatomic
distance of 269 pm, while the Ca–Oa bond is already broken
(Ca–Oa = 250 pm). Herein, the [C4H9] fragment is stabilized via
hydrogen bonds formed by a-hydrogen atoms (H1 and H2) with
the oxygen atoms of the physisorbed alcohol (O2) and the zeolite Fig. 8. Gibbs free energy profile for the dehydration of 1-butanol to dibutyl ether
framework (Oa) respectively. The results of the Bader charge (DBE) in H-ZSM-5 at 400 K.
38 M. John et al. / Journal of Catalysis 330 (2015) 28–45

respect to the gas phase butanol and the zeolite. The energetic leads to the formation of a co-adsorbed species C4, having both
preference for the dimer mediated route over the alkoxide medi- butanol and butene adsorbed on the acid site of the zeolite.
ated route for ether formation reaction, has also been reported Further desorption of butanol (step 21) and butene (step 8)
for dehydration of methanol [66,67] and ethanol [68] in zeolites. restores the Brønsted acid site.
The free energy diagram suggests that the conversion of monomer In the transition state (TS-10, Fig. 9) structure, the b-hydrogen
M2 to 1-butoxide (via SN2 substitution reaction, TS-4) having the (H4) and the leaving group (butanol) have a near syn-coplanar con-
highest apparent free energy barrier, would be critical for the figuration with a O2–Ca–Cb–H4 dihedral angle of 21.6°. The
butoxide-mediated conversion of alcohol to ether (mechanisms 7 a-carbon is completely planar, while the b-carbon is approaching
and 8). For the further conversion of 1-butoxide to ether, the SN2 a planar state (see supporting information, Table S3). The extent
type nucleophilic substitution (via TS-8) offers a much lower free of the Ca–O2 bond scission is larger than that of the Cb–H4 bond,
energy barrier as compared to the SN1 like nucleophilic substitu- giving the syn-elimination an E1 character. Moreover, the inter-
tion (via TS-9). The protonated ether molecule that is formed can atomic distances show that the Ca–O2 (229.9 pm) and Cb–H4
either undergo further reaction (see path C) or get deprotonated (131.1 pm) bonds are nearly broken, which is indicative of a late
and desorb. Although the desorption of ether is highly endothermic transition state.
(DHo = 191 kJ/mol), the overall conversion of two molecules of The physisorbed butanol molecule has a minor charge of
1-butanol to ether and water is exothermic and hence, should be +0.1026, while the zeolite framework has a charge of 0.7491.
favored at low temperatures. The charge on the [C4H9] fragment is +0.6465, but the major frac-
tion of this charge is associated with the b-hydrogen (+0.4113)
3.2.3. Path C: Decomposition of di-1-butyl ether to 1-butene which has nearly broken its bond with the b-carbon (Cb). Based
The experimental study by Makarova et al. [17] indicated a on the geometrical aspects and the results of the Bader charge
decrease in ether yield with increasing site time, which was cred- analysis, one can conclude that the key step of this reaction mech-
ited to the ether decomposition reaction. Moreover, the observa- anism follows a one step 1,2-syn-elimination reaction with E1
tion of similar reaction rates for dehydration of butanol to character.
butene and for ether decomposition to butene indicates that
butene formation can proceed via an ether-mediated route. In view 3.2.3.2. Reaction mechanism 10 (decomposition of ether to butene and
of this, two plausible mechanisms for the decomposition of ether butanol via E2 (anti) elimination). This reaction mechanism is sim-
to butene have been studied. Herein, the ether decomposition ilar to mechanism 9, but proceeds via an anti-elimination route
mechanism involves either a syn- or an anti- (E2) elimination reac- instead of the syn-elimination. The protonated ether undergoes a
tion as described in the case of direct dehydration of butanol to sequence of elementary steps, namely reorientation (step 22) to
butene. The transition state structures associated with these two an antiperiplanar configuration DBE2, butanol elimination (step
mechanisms are shown in Fig. 9. 23) to form physisorbed 1-butene and desorption of 1-butene (step
8) to restore the zeolite Brønsted acid site. The elementary step 23
3.2.3.1. Reaction mechanism 9 (decomposition of ether to butene and is activated (TS-11) and has an activation energy of 84 and
butanol via syn elimination). This reaction mechanism involves the 147 kJ/mol, considering DBE2 (Table 3) and DBE⁄ as reference
decomposition of the protonated ether produced via the states, respectively. In this step, the protonated ether molecule
re-adsorption of dibutyl ether (reverse of step 16), in a sequence undergoes an elimination through a concerted mechanism, involv-
of elementary steps 20, 21 and 8. In step 20, the protonated ether ing simultaneous cleavage of the Ca–O2 bond and abstraction of
undergoes a cleavage of a C–O bond and a concurrent abstraction the b-hydrogen (H4) by the zeolite basic oxygen (Oa).
of a b-hydrogen by the zeolite basic oxygen. This elementary step In the transition state (TS-11, Fig. 9) geometry, the b-hydrogen
is activated (TS-10) and has an activation energy of 140 kJ/mol (H4) points toward the bridge oxygen atom (Oa) and the leaving
(Table 3), considering DBE⁄ as reference state. This is in good agree- group (butanol) is positioned opposite to it, having a dihedral angle
ment with the experimentally reported value of 138 ± 8 kJ/mol for (O2–Ca–Cb–H4) of 181.4°. The a-carbon and b-carbon have a near
decomposition of ether to butene [17]. The elementary step 20 planar configuration with the a-carbon being closer to planarity

Fig. 9. Transition state structures (views at the channel intersection) for the ether decomposition reaction (path C): (1) TS-10, syn-elimination (mechanism 9), (2) TS-11, anti-
elimination (mechanism 10). Color code: silicon – light blue, aluminum – pink, oxygen – red, hydrogen – white, carbon – gray, hydrogen bonds (distance < 250 pm) – green
dashed lines, bond breaking/forming – black dashed lines. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)
M. John et al. / Journal of Catalysis 330 (2015) 28–45 39

than the b-carbon (see bond angles in supporting information, and Table 2) has been considered without making any assumption
Table S3). The Ca–O2 and Cb–H4 bond lengths in the transition state for the rate determining step. The effect of reaction conditions, viz.,
are 222 and 135 pm respectively. The extent of C–O bond breakage partial pressure of 1-butanol (0.01–100 kPa) and water (0–40 kPa),
is more pronounced than that of the b-carbon (Cb) and b-hydrogen site time (NH+/FBuOH,0 = 0–200 mol H+ s/mol BuOH0) and reaction
(H4). temperature (400–460 K), is studied.
The physisorbed butanol molecule has a minor charge of A comparison of the forward, reverse and net reaction rates for
+0.1182, while the zeolite framework has a charge of 0.812. the different elementary steps over a wide range of reaction condi-
The charge on the [C4H9] fragment is +0.6938, but the major frac- tions has been used to determine the rate determining and the
tion of this charge is associated with the b-hydrogen (+0.3824) equilibrated/non-equilibrated steps along each mechanism. For
which has nearly broken its bond with the b-carbon (Cb). Based the direct conversion of butanol to butene (path A), the elementary
on the geometrical aspects and the results of the Bader charge steps R2 (TS-1), R4 (TS-2), R7 (TS-3) and R13 (TS-6) are found to be
analysis, one can conclude that the key step of this reaction mech- the rate determining steps along mechanism 1 (via E1-like
anism follows an 1,2-anti-elimination reaction. elimination), mechanism 2 (via syn-elimination), mechanism 3
(via anti-elimination) and mechanism 5 (butanol-assisted
3.2.3.3. Comparison of the reaction mechanisms for the decomposition syn-elimination), respectively. On the other hand, mechanism 4
of ether. The Gibbs free energy profile for the two plausible mech- (butoxide-mediated dehydration of butanol to butene) involves
anisms of the ether decomposition to 1-butene (path C) is shown in two non-equilibrated steps, namely step R9 (TS-4) and R10
Fig. 10. Among the syn- (TS-10) and anti-elimination (TS-11), the (TS-5). For the conversion of butanol to ether (path B), the elemen-
syn-elimination offered a lower free energy barrier (apparent tary step R15 (TS-7) is the rate determining step for mechanism 6
DGà of 4 kJ/mol at 400 K with respect to the gas phase butanol (butanol dimer to ether via SN2 reaction), while for both the
and the zeolite). butoxide-mediated etherification mechanisms (7 via SN2 and 8
via SN1), two non-equilibrated steps (steps R9 (TS-4) and R18
3.2.4. Comparison of the different reaction pathways for 1-butanol (TS-8) for mechanism 7 and R9 (TS-4) and R19 (TS-9) for
dehydration mechanism 8) are identified. For the ether decomposition reaction
Comparison of the free energy profiles for paths A, B, and C, (path C), elementary steps R20 (TS-10) and R23 (TS-11) are found to
indicates that etherification (via TS7, SN2 substitution, see Fig. 8) be the rate determining steps along mechanism 9 (via
and ether decomposition (via TS-10, syn-elimination, see Fig. 10) syn-elimination) and 10 (via anti-elimination), respectively. All
to be energetically more favorable as compared to any of the mech- other reaction steps in the microkinetic model are found to be
anisms associated with path A (see Fig. 6). However, as the actual quasi-equilibrated.
reaction rate is a complex function of reaction conditions and feed To validate the results of our ab initio based microkinetic
concentrations, it is imperative to carry out a detailed reaction model, a comparison is made with the experimental result
path analysis to arrive at a conclusive result regarding the domi- reported by Makarova et al. [17]. The microkinetic model was used
nant reaction mechanism in actual reaction conditions. to simulate their experimental conditions and the turnover fre-
quency (TOF) for the production of dibutyl ether and butene was
compared at identical site times. The experimentally reported
3.3. Microkinetic model and reaction path analysis
reaction rates were converted to TOF by accounting for the number
of Brønsted acid sites per gram of catalyst (see supporting informa-
A detailed microkinetic model involving 10 different reaction
tion for details). Table 4 provides a comparison between our theo-
mechanisms, consisting of 23 elementary steps (as seen in Fig. 4
retical and literature reported experimental results.
The ab initio based microkinetic model captures the experimen-
tal observation of a considerably larger ether formation compared
to butene formation under the investigated conditions. The theo-
retical results predict reasonably well the rate of production of
ether but under-predict the rate of formation of butene by an order
of magnitude. This can be attributed to deficiencies of GGA func-
tionals (e.g. overpolarization effects [49]), inaccuracies in describ-
ing dispersive interactions [40–43] and the inaccuracy associated
with the harmonic oscillator approximation in predicting the
entropic contributions for the gas-phase butene molecule [69]
and loosely bonded transition states [70]. Nevertheless, this devia-
tion from the experimental result for butene formation is in the
order of the expected accuracy for the harmonic oscillator approx-
imation [69,70]. Moreover, the total rate of butanol dehydration is
Fig. 10. Gibbs free energy profile for the decomposition of ether to 1-butene in comparable as described by the similar conversion values. Overall,
H-ZSM-5 at 400 K (Gibbs free energy of zeolite with two gas-phase 1-butanol the simulated results are in good agreement with the experimen-
molecules was considered as the reference zero). tally observed results and the model can be used to identify

Table 4
Comparison between theoretical and experimental TOFs for production of dibutyl ether (DBE) and butene at 400 K, butanol feed mole fraction 0.7%, total pressure of 1 bar and site
time of 36.6 molH+ s mol1.

This work Experimentala


+ 6
1 TOF for production of butene (mol/mol H /s) 2.5  10 4.1  105
2 TOF for production of DBE (mol/mol H+/s) 1.7  104 5.1  104
3 Conversion (mol%) 1.8 2
a
Experimental result of Makarova et al. [17].
40 M. John et al. / Journal of Catalysis 330 (2015) 28–45

dominant reaction mechanisms and to provide a reliable insight predominantly occurs via decomposition of ether (path C), except
into the effect of reaction conditions. for very low conversions, where it occurs via direct dehydration
of butene (path A). This can be explained on the basis of the change
in the surface coverages with an increase in conversion (see
3.3.1. Effect of conversion
Fig. 11d). At a very low conversion, the adsorbed dimer (D1) is
The influence of conversion on product yields, TOFs and surface
the most abundant species, followed by protonated monomer
coverages is studied by systematically increasing the site time at
(M1) and protonated ether (DBE⁄), and the production of butene
constant temperature and pressure, thus increasing the conversion
occurs via path A. With an increase in conversion, DBE⁄ becomes
of butanol. The butene yield was found to increase steadily with
the most abundant surface species and butene is produced via
increasing site time, while the ether yield was found to pass
ether decomposition.
through a maximum as a function of site time (see Fig. 11a). The
The simulated surface coverages (see Fig. 11d) indicate a large
simulated results are in line with the experimentally observed
abundance of adsorbed dibutyl ether (DBE⁄) and dimeric species
trends for the butene and ether yield [17]. The decrease in ether
(D1) over the zeolite active sites. TGA studies [71] for 1-butanol
yield at higher site time is attributed to the decomposition of ether
adsorption on H-ZSM-5 indicated an occupancy of two molecules
to butene.
of alcohol per Al site. In addition, IR studies [72] for adsorption
Fig. 11c shows the effect of conversion on the TOFs for the dif-
of 1-butanol on H-ZSM-5 indicated that the surface dominant spe-
ferent reaction pathways. Here, the TOF value for each reaction
cies have a stoichiometry close to that of DBE.
pathway is calculated by summing up the TOF values for each of
the contributing reaction mechanism. The detailed contribution
of the TOF values of each reaction mechanism to the corresponding 3.3.2. Effect of reaction temperature
reaction pathway at 450 K and 1 kPa feed butanol pressure is The effect of temperature on product selectivity, surface cov-
shown in Fig. S6 of the supporting information. Reaction path A erage and TOFs is investigated at a temperature range of 400–
proceeds predominantly via reaction mechanism 3 (E2 elimina- 460 K. An increase in the reaction temperature is also associated
tion) and mechanism 5 (butanol-assisted syn-elimination), while with an increase in the conversion which has a significant
reaction paths B and C occur primarily via mechanism 6 (SN2 sub- impact on the overall product selectivity. Therefore, when com-
stitution of butanol dimer to ether) and mechanism 9 paring selectivities at different temperatures, the conversion
(syn-elimination), respectively. Overall, it is seen that reaction path needs to be fixed in order to decouple the temperature effect
B remains the dominant one up to 10% conversion, while path C from the conversion effect. The temperature of the study was
starts gaining importance with an increase in conversion. limited to 460 K in order to avoid reaction conditions which
Butene selectivity is found to increase with increasing conver- would lead to further conversion of butene to higher hydrocar-
sion (see Fig. 11b). As seen in Fig. 11c, the production of butene bon and cracked products.

Fig. 11. (a) Butanol conversion and product (1-butene and DBE) molar yield as a function of site time, (b) 1-butene and DBE selectivity as a function of conversion, (c)
turnover frequencies (TOF) for different reaction pathways as a function of conversion, (d) coverages for surface species. Reaction conditions: 450 K, butanol feed partial
pressure of 1 kPa and site time varied between 0 and 30 (NH+(mol))/FBuOH,0(mol/s)).
M. John et al. / Journal of Catalysis 330 (2015) 28–45 41

Fig. 12. Effect of reaction temperature on (a) 1-butene selectivity and (b) DBE selectivity as a function of conversion. (c) Turnover frequencies (TOF) for different reaction
pathways at x = 10% and (d) surface coverages compared at a constant butanol conversion (x) of 10%. Inlet partial pressure of butanol: 1 kPa, site time adjusted in the range of
0–200 (NH+(mol))/FBuOH,0(mol/s)) so as to attain similar conversion levels.

Fig. 12a and b shows the effect of temperature variation on pro- butanol flow rate and varying the inert gas flow rate to match
duct selectivity at constant inlet pressure (1 kPa) of butanol. At low the desired inlet partial pressure value. The effect of inlet butanol
temperatures (below 400 K), ether remains the key product in partial pressures on the product selectivity is compared at different
agreement with the low temperature (368–409 K) experimental conversion levels (Fig. 13a and b).
results reported by Chiang and Bhan [68]. On the other hand, selec- An increase in butanol feed partial pressure leads to a decrease
tivity shifts toward the production of butene with an increase in in butene selectivity (see Fig. 13a). This is attributed to the
temperature, which is consistent with high temperature (673– decrease in the TOF of path A and path C (see Fig. 13c). This is in
773 K) experimental results where formation of an ether fraction turn ascribed to the increased preference for adsorption of butanol
is not observed [18,19]. to form the adsorbed dimer (as compared to protonated ether or
Fig. 12c shows the effect of temperature on TOFs compared at a adsorbed butanol monomer) at higher butanol partial pressures
conversion level of 10%. In the temperature window of 400–460 K, (see Fig. 13d).
the rate of production of butene occurs essentially via an Fig. 13c depicts a shift in dominant path with inlet butanol pres-
ether-mediated consecutive reaction scheme (i.e. path B followed sure. For extremely low values of inlet butanol pressure (PBuOH,0
by path C). This is consistent with the conclusion drawn on the less than 0.01 kPa) path A is dominant. With an increase in pres-
basis of comparison of the free energy diagrams for the different sure (PBuOH,0 in the range of 0.1–10 kPa), butanol dehydration
reaction pathways (Section 3.2.4). essentially proceeds via paths B and C. For pressures greater than
Finally, temperature plays a key role in defining the dominant 10 kPa, path B becomes the dominant reaction pathway. Hence,
reaction mechanism within each reaction pathway. The detailed one should be careful when comparing results obtained from low
contribution of the TOF values of each reaction mechanism to the pressure experiments (such as FTIR or Temporal Analysis of
corresponding reaction pathway is shown in Fig. S7 of the support- Products) to that of the high pressure experimentation carried
ing information. For path A, there is a shift in the dominant mech- out in a micro-reactor or an industrial scale reactor.
anism from mechanism 5 (butanol-assisted syn-elimination) to Further insight into the pressure dependence of TOFs for each
mechanism 3 (E2 elimination) with an increase in temperature. reaction pathway is obtained by looking into the TOFs for the indi-
On the other hand, reaction mechanisms 6 (SN2 butanol dimer to vidual reaction mechanisms (see Fig. S8 of the supporting informa-
ether) and 9 (syn-elimination) remain dominant for paths B and tion). For path A, the increase in butanol partial pressure is
C, respectively. associated with a decrease in the TOFs for mechanisms 1–4 and
an increase in the TOF for mechanism 5. This leads to a shift in
3.3.3. Effect of partial pressure of 1-butanol the dominant mechanism from E2 elimination (mechanism 3) to
The effect of butanol partial pressure on the butene and ether butanol-assisted syn-elimination (mechanism 5) with an increase
selectivity has been examined using microkinetic simulations in in butanol partial pressure. For path B, the increase in butanol par-
the butanol inlet partial pressures range of 0.001–100 kPa at tial pressure is associated with an increase in the TOF for mecha-
450 K. The butanol partial pressure is varied by keeping the same nism 6 and a decrease in the TOFs for mechanisms 7 and 8. The
42 M. John et al. / Journal of Catalysis 330 (2015) 28–45

Fig. 13. Effect of butanol partial pressure in the feed on (a) 1-butene selectivity and (b) DBE selectivity as a function of conversion. (c) Turnover frequencies (TOF) for different
reaction pathways at x = 10% and (d) surface coverages compared at a constant butanol conversion (x) of 10%. Reaction temperature: 450 K, site time is varied between 0 and
30 (NH+(mol))/FBuOH,0(mol/s)) in order to attain similar conversion levels.

reaction mechanism 6 (SN2, butanol dimer to ether) remains dom- 3.3.4. Effect of partial pressure of water
inant except at very low partial pressures (less than 0.001 kPa), The effect of the partial pressure of water on the butanol dehy-
where mechanism 7 (SN2 butoxide-mediated etherification) tends dration reaction is of prime interest for the production of alkenes
to gain importance. For path C, both mechanisms 9 and 10 depict a from bio-butanol. The fermentation process used for production
zero-order pressure dependence until butanol feed partial pressure of bio-butanol, utilizes water as a solvent and involves production
of 1 kPa, while shifting to a negative order pressure dependence at of acetone and ethanol in addition to 1-butanol. Typically, an
higher butanol partial pressures. Thus, for experimentally relevant azeotropic distillation process [73] is used to recover an aqueous
conditions (PBuOH,0 above 1 kPa) one would observe a negative butanol stream from the clarified fermentation broth. The
order dependence for the formation of butene with respect to buta- 1-butanol–water mixture has an azeotropic composition of 24.8
nol partial pressure. Reaction mechanism 9, i.e. ether decomposi- and 75.2 mol% of butanol and water, respectively. Since further
tion via syn-elimination, remains dominant throughout the concentration of 1-butanol is energy intensive, it would be ideal
complete pressure range for path C. to feed the water–butanol azeotrope for the dehydration process
The pressure dependence and shift in the dominant reaction [74].
path can be explained on the basis of changes in the surface cover- Inhibition effects due to the presence of water can be ascribed
age with reaction conditions. The simulated surface coverages (see to several different factors such as competitive adsorption and
Fig. 13d) depict a large abundance of adsorbed dibutyl ether (DBE⁄) co-adsorption, solvation effect by water (where water stabilizes
and dimeric species (D1) over the zeolite active sites especially at the reactant better than the transition state structure), or
high pressures. Since the surface coverages of D1 and DBE⁄ are hydrothermal deactivation due to change in catalyst structure
much higher than M1 at high butanol pressures, this leads to (for instance hydrothermal dealumination in zeolites). Moreover,
prevalence of path B and path C mechanisms at these conditions. at higher water partial pressure thermodynamic limitations can
On the other hand, the M1 intermediate is found to have a signif- also play a role. The present study essentially focuses on the com-
icant surface coverage only at low partial pressures of butanol, petitive adsorption and co-adsorption of a water molecule.
where path A prevails. In order to investigate the effect of water on butanol dehydra-
Although the present study infers that the alcohol dimer can tion within H-ZSM-5, a wide range of inlet water partial pressures
undergo further reaction, it is pertinent to note that the formation (PH2O,0 – 0.1–40 kPa) has been studied at a constant inlet butanol
of these dimer species reduces the overall rate of production of partial pressure of 1 kPa and reaction temperature of 450 K as
alkenes (see Fig. 13c/d). Our simulations indicate a decrease in shown in Fig. 14. Interestingly, the presence of water in the feed
TOFs for path A and path C which leads to the formation of butene does not have any significant impact on the butanol conversion
with an increase in butanol partial pressure (associated with and the overall product selectivity even at partial pressure ratios
increase in dimer coverage). A decrease in ethylene yield was also much higher than that corresponding to the azeotropic composi-
observed in ethanol dehydration over H-MOR [68], with an tion (PH2O,0/PBuOH,0|azeotrope  3). This is also evident from the neg-
increase in ethanol pressure. ligible change in the TOFs (see Fig. 14a) with an increase in
M. John et al. / Journal of Catalysis 330 (2015) 28–45 43

Fig. 14. Effect of partial pressure of water on (a) turnover frequencies (TOF) for different reaction pathways at x = 10% and (b) surface coverages compared at a constant
butanol conversion (x) of 10%. Reaction temperature: 450 K, inlet partial pressure of butanol: 1 kPa, site time adjusted in the range of 0–30 (NH+(mol))/FBuOH,0(mol/s)) so as to
attain similar conversion levels.

partial pressure of water indicative of a zero order dependence dehydration of butanol to butene (path A), mechanisms involving
with respect to water. Although this seems to be in contrast to an E2 (anti) elimination reaction remain dominant at low butanol
the experimental observation for the dehydration of 2-butanol partial pressures (below 0.01 kPa) and shift to butanol-assisted
over POM cluster [20] and ethanol dehydration over c-alumina syn-elimination at higher butanol pressures (greater than 1 kPa).
[75], it has been reported in the literature [76,77] that water has The ether formation (path B) can occur via SN2 mechanisms fol-
no inhibiting effect on the dehydration of ethanol in H-ZSM-5. lowing two possible routes, from physisorbed dimer (D2) or from
The observation that the partial pressure of water has practically 1-butanol and 1-butoxide. Ether formation predominantly occurs
no effect on the rates can be explained by the marginal increase via the former route, except at very low partial pressures where
of the water coverage with increasing partial pressure of water it can occur via both routes. The ether decomposition (path C)
(see Fig. 14b) which is related to the higher adsorption strength occurs primarily via a 1,2-syn-elimination. The shift in the
for the butanol dimer (D1) and ether (DBE) in H-ZSM-5 as com- dominant reaction mechanism can be explained on the basis of
pared to the water (W) and the co-adsorbed butanol–water species changes in surface coverages. This shift in the preferred reaction
(C2). A similar reasoning was proposed by Corma and mechanism can help to reconcile conflicting observations reported
Perezpariente [78] to explain the absence of a water inhibition at different reaction conditions. Last but not the least, the absence
effect for the ethanol dehydration reaction over an acidic sepiolite of a water inhibition effect makes the dehydration of aqueous
catalyst. The higher adsorption strength of 1-butanol in H-ZSM-5 bio-butanol an attractive option for the production of
as compared to water and other lower alcohols (methanol/etha bio-butenes, which can serve as a feedstock for the chemical
nol/1-propanol/2-propanol) is also consistent with the experimen- industry.
tal observation reported by Aronson et al. [79] that lower alcohols
were unable to displace adsorbed butanol in the zeolite. Acknowledgments

4. Conclusions This work is supported by the Long Term Structural


Methusalem Funding by the Flemish Government – Grant No.
This study provides a detailed insight into the plausible reaction BOF09/01M00409. The computational resources (Stevin
mechanisms for dehydration of 1-butanol in H-ZSM-5 zeolite. A Supercomputer Infrastructure) and services used in this work were
first principles based microkinetic model is used to obtain a predic- provided by Ghent University.
tive guidance on the effect of reaction conditions on reaction rates
and product selectivity. Useful insights into the dominant reaction Appendix A. Supplementary material
mechanism and its dependence on reaction conditions are derived
using a reaction path analysis. Supplementary data associated with this article can be found, in
Our study reveals the importance of the alcohol protonation by the online version, at http://dx.doi.org/10.1016/j.jcat.2015.07.005.
the Brønsted acid site, which helps in the advancement of the reac-
tion as it leads to an elongation of the C–O bond, facilitating the
References
cleavage of C–O bond in the later steps. The reasonable agreement
between the simulated and literature reported experimental val- [1] R. Luque, L. Herrero-Davila, J.M. Campelo, J.H. Clark, J.M. Hidalgo, D. Luna, J.M.
ues, demonstrates the reliability of the DFT based microkinetic Marinas, A.A. Romero, Biofuels: a technological perspective, Energy Environ.
model in providing directional information on reaction kinetics. Sci. 1 (2008) 542–564.
[2] M.Z. Jacobson, Effects of ethanol (E85) versus gasoline vehicles on cancer and
An increase in conversion leads to an increase in butene yield. mortality in the United States, Environ. Sci. Technol. 41 (2007) 4150–4157.
Lower temperatures favor ether formation, while higher tempera- [3] R. Cook, S. Phillips, M. Houyoux, P. Dolwick, R. Mason, C. Yanca, M. Zawacki, K.
tures lead to preferential production of butene. Higher butanol par- Davidson, H. Michaels, C. Harvey, J. Somers, D. Luecken, Air quality impacts of
increased use of ethanol under the United States’ Energy Independence and
tial pressures favor dehydration of butanol to ether, followed by a
Security Act, Atmos. Environ. 45 (2011) 7714–7724.
temperature dependent ether decomposition. The reaction path [4] N. Savage, The ideal biofuel, Nature 474 (2011) S9–S11.
analysis reveals that, except for extremely low butanol pressures, [5] J.Q. Guan, C. Xu, B. Liu, Y. Yang, Y.Y. Ma, Q.B. Kan, Partial oxidation of isobutane
the production of butene occurs essentially via an ether mediated over hydrothermally synthesized Mo–V–Te–O mixed oxide catalysts, Catal.
Lett. 126 (2008) 301–307.
route. The butanol partial pressure has a prominent role in deter- [6] A.A. Gronowski, Synthesis of butyl rubber in hexane using a mixture of Et2AlCl
mining the dominant reaction mechanism. For the direct and EtAlCl2 in the initiating system, J. Appl. Polym. Sci. 87 (2003) 2360–2364.
44 M. John et al. / Journal of Catalysis 330 (2015) 28–45

[7] G. Busca, Acid catalysts in industrial hydrocarbon chemistry, Chem. Rev. 107 [37] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh, C.
(2007) 5366–5410. Fiolhais, Atoms, molecules, solids, and surfaces – applications of the
[8] M.E. Wright, B.G. Harvey, R.L. Quintana, Highly efficient zirconium-catalyzed generalized gradient approximation for exchange and correlation, Phys. Rev.
batch conversion of 1-butene: a new route to jet fuels, Energy Fuels 22 (2008) B 46 (1992) 6671–6687.
3299–3302. [38] S. Grimme, Semiempirical GGA-type density functional constructed with a
[9] O. Olaofe, P.L. Yue, Mechanism of the dehydration of 1-butanol over zeolites, long-range dispersion correction, J. Comput. Chem. 27 (2006) 1787–1799.
Collect. Czech. Chem. C 50 (1985) 1834–1841. [39] T. Kerber, M. Sierka, J. Sauer, Application of semiempirical long-range
[10] C. Williams, M.A. Makarova, L.V. Malysheva, E.A. Paukshtis, K.I. Zamaraev, J.M. dispersion corrections to periodic systems in density functional theory, J.
Thomas, Mechanistic studies of the catalytic dehydration of isobutyl alcohol Comput. Chem. 29 (2008) 2088–2097.
on NaH-ZSM-5, J. Chem. Soc. Faraday Trans. 86 (1990) 3473–3485. [40] C. Tuma, J. Sauer, Treating dispersion effects in extended systems by hybrid
[11] C. Williams, M.A. Makarova, L.V. Malysheva, E.A. Paukshtis, E.P. Talsi, J.M. MP2: DFT calculations – protonation of isobutene in zeolite ferrierite, Phys.
Thomas, K.I. Zamaraev, Kinetic-studies of catalytic dehydration of tert-butanol Chem. Chem. Phys. 8 (2006) 3955–3965.
on zeolite NaH-ZSM-5, J. Catal. 127 (1991) 377–392. [41] F. Goltl, A. Gruneis, T. Bucko, J. Hafner, Van der Waals interactions between
[12] M.A. Makarova, C. Williams, K.I. Zamaraev, J.M. Thomas, Mechanistic study of hydrocarbon molecules and zeolites: periodic calculations at different levels of
sec-butyl alcohol dehydration on zeolite H-ZSM-5 and amorphous theory, from density functional theory to the random phase approximation
aluminosilicate, J. Chem. Soc. Faraday Trans. 90 (1994) 2147–2153. and Moller–Plesset perturbation theory, J. Chem. Phys. 137 (2012).
[13] M.A. Makarova, C. Williams, J.M. Thomas, K.I. Zamaraev, Dehydration of N- [42] M.V. Vener, X. Rozanska, J. Sauer, Protonation of water clusters in the cavities
butanol on HNa-ZSM-5, Catal. Lett. 4 (1990) 261–264. of acidic zeolites: (H2O)(n)H-chabazite, n = 1–4, Phys. Chem. Chem. Phys. 11
[14] M.A. Makarova, C. Williams, V.N. Romannikov, K.I. Zamaraev, J.M. Thomas, (2009) 1702–1712.
Influence of pore confinement on the catalytic dehydration of isobutyl alcohol [43] C.C. Chiu, G.N. Vayssilov, A. Genest, A. Borgna, N. Rosch, Predicting adsorption
on H-ZSM-5, J. Chem. Soc. Faraday Trans. 86 (1990) 581–584. enthalpies on silicalite and HZSM-5: a benchmark study on DFT strategies
[15] M.A. Makarova, E.A. Paukshtis, J.M. Thomas, C. Williams, K.I. Zamaraev, In situ addressing dispersion interactions, J. Comput. Chem. 35 (2014) 809–819.
FTIR kinetic-studies of diffusion, adsorption and dehydration reaction of tert- [44] C.M. Nguyen, M.F. Reyniers, G.B. Marin, Adsorption thermodynamics of C1–C4
butanol on zeolite NaH-ZSM-5, Catal. Today 9 (1991) 61–68. alcohols in H-FAU, H-MOR, H-ZSM-5, and H-ZSM-22, J. Catal. 322 (2015) 91–
[16] M.A. Makarova, E.A. Paukshtis, J.M. Thomas, C. Williams, K.I. Zamaraev, In-situ 103.
FTIR and GC kinetic-studies – complementary methods in the mechanistic [45] C.M. Nguyen, M.F. Reyniers, G.B. Marin, Theoretical study of the adsorption of
study of butanol dehydration on zeolite H-ZSM-5, Stud. Surf. Sci. Catal. 75 C1–C4 primary alcohols in H-ZSM-5, Phys. Chem. Chem. Phys. 12 (2010)
(1993) 1711–1714. 9481–9493.
[17] M.A. Makarova, E.A. Paukshtis, J.M. Thomas, C. Williams, K.I. Zamaraev, [46] C.M. Nguyen, M.F. Reyniers, G.B. Marin, Theoretical study of the adsorption of
Dehydration of N-butanol on zeolite H-ZSM-5 and amorphous aluminosilicate the butanol isomers in H-ZSM-5, J. Phys. Chem. C 115 (2011) 8658–8669.
– detailed mechanistic study and the effect of pore confinement, J. Catal. 149 [47] T. Bučko, J. Hafner, The role of spatial constraints and entropy in the adsorption
(1994) 36–51. and transformation of hydrocarbons catalyzed by zeolites, J. Catal. 329 (2015)
[18] D.Z. Zhang, R. Al-Hajri, S.A.I. Barri, D. Chadwick, One-step dehydration and 32–48.
isomerisation of n-butanol to iso-butene over zeolite catalysts, Chem. [48] F. Leydier, C. Chizallet, D. Costa, P. Raybaud, Revisiting carbenium chemistry
Commun. 46 (2010) 4088–4090. on amorphous silica–alumina: unraveling their milder acidity as compared to
[19] D.Z. Zhang, S.A.I. Barri, D. Chadwick, n-Butanol to iso-butene in one-step over zeolites, J. Catal. 325 (2015) 35–47.
zeolite catalysts, Appl. Catal. A – Gen. 403 (2011) 1–11. [49] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, A consistent and accurate ab initio
[20] J. Macht, M.J. Janik, M. Neurock, E. Iglesia, Mechanistic consequences of parametrization of density functional dispersion correction (DFT-D) for the 94
composition in acid catalysis by polyoxometalate Keggin clusters, J. Am. Chem. elements H–Pu, J. Chem. Phys. 132 (2010).
Soc. 130 (2008) 10369–10379. [50] R.F.W. Bader, Atoms in Molecules: A Quantum Theory, Clarendon Press,
[21] J. Macht, M.J. Janik, M. Neurock, E. Iglesia, Catalytic consequences of Oxford, 1990.
composition in polyoxometalate clusters with Keggin structure, Angew. [51] G. Henkelman, A. Arnaldsson, H. Jonsson, A fast and robust algorithm for
Chem. Int. Ed. 46 (2007) 7864–7868. Bader decomposition of charge density, Comput. Mater. Sci. 36 (2006)
[22] J. Macht, R.T. Carr, E. Iglesia, Consequences of acid strength for isomerization 354–360.
and elimination catalysis on solid acids, J. Am. Chem. Soc. 131 (2009) 6554– [52] G. Henkelman, H. Jonsson, Improved tangent estimate in the nudged elastic
6565. band method for finding minimum energy paths and saddle points, J. Chem.
[23] M.J. Janik, J. Macht, E. Iglesia, M. Neurock, Correlating acid properties and Phys. 113 (2000) 9978–9985.
catalytic function: a first-principles analysis of alcohol dehydration pathways [53] G. Henkelman, H. Jonsson, A dimer method for finding saddle points on high
on polyoxometalates, J. Phys. Chem. C 113 (2009) 1872–1885. dimensional potential surfaces using only first derivatives, J. Chem. Phys. 111
[24] J. Macht, R.T. Carr, E. Iglesia, Functional assessment of the strength of solid acid (1999) 7010–7022.
catalysts, J. Catal. 264 (2009) 54–66. [54] A. Heyden, A.T. Bell, F.J. Keil, Efficient methods for finding transition states in
[25] A. Prestianni, R. Cortese, D. Duca, Propan-2-ol dehydration on H-ZSM-5 and H– chemical reactions: comparison of improved dimer method and partitioned
Y zeolite: a DFT study, React. Kinet. Mech. Catal. 108 (2013) 565–582. rational function optimization method, J. Chem. Phys. 123 (2005).
[26] D.H. Olson, N. Khosrovani, A.W. Peters, B.H. Toby, Crystal structure of [55] B.A. De Moor, A. Ghysels, M.F. Reyniers, V. Van Speybroeck, M. Waroquier, G.B.
dehydrated CsZSM-5 (5.8Al): evidence for nonrandom aluminum Marin, Normal mode analysis in zeolites: toward an efficient calculation of
distribution, J. Phys. Chem. B 104 (2000) 4844–4848. adsorption entropies, J. Chem. Theory Comput. 7 (2011) 1090–1101.
[27] J. Dedecek, V. Balgova, V. Pashkova, P. Klein, B. Wichterlova, Synthesis of ZSM- [56] Y. Zhao, D.G. Truhlar, Computational characterization and modeling of
5 zeolites with defined distribution of Al atoms in the framework and buckyball tweezers: density functional study of concave–convex pi center
multinuclear MAS NMR analysis of the control of Al distribution, Chem. Mater. dot center dot center dot pi interactions, Phys. Chem. Chem. Phys. 10 (2008)
24 (2012) 3231–3239. 2813–2818.
[28] S.R. Lonsinger, A.K. Chakraborty, D.N. Theodorou, A.T. Bell, The effects of local [57] R.F. Ribeiro, A.V. Marenich, C.J. Cramer, D.G. Truhlar, Use of solution-phase
structural relaxation on aluminum siting within H-ZSM-5, Catal. Lett. 11 vibrational frequencies in continuum models for the free energy of solvation, J.
(1991) 209–217. Phys. Chem. B 115 (2011) 14556–14562.
[29] B.F. Mentzen, M. Sacerdoteperonnet, Prediction of preferred proton locations [58] S. Grimme, Supramolecular binding thermodynamics by dispersion-corrected
in HMFI benzene complexes by molecular mechanics calculations – density functional theory, Chem. – Eur. J. 18 (2012) 9955–9964.
comparison with NMR, structural and calorimetric results, Mater. Res. Bull. [59] J.H. Jensen, Predicting accurate absolute binding energies in aqueous solution:
29 (1994) 1341–1348. thermodynamic considerations for electronic structure methods, Phys. Chem.
[30] S. Svelle, C. Tuma, X. Rozanska, T. Kerber, J. Sauer, Quantum chemical modeling Chem. Phys. 17 (2015) 12441–12451.
of zeolite-catalyzed methylation reactions: toward chemical accuracy for [60] G. Piccini, J. Sauer, Effect of anharmonicity on adsorption thermodynamics, J.
barriers, J. Am. Chem. Soc. 131 (2009) 816–825. Chem. Theory Comput. 10 (2014) 2479–2487.
[31] G. Kresse, J. Hafner, Abinitio molecular-dynamics for liquid-metals, Phys. Rev. [61] G. Piccini, M. Alessio, J. Sauer, Y.C. Zhi, Y. Liu, R. Kolvenbach, A. Jentys, J.A.
B 47 (1993) 558–561. Lercher, Accurate adsorption thermodynamics of small alkanes in zeolites. Ab
[32] G. Kresse, J. Hafner, Ab-initio molecular-dynamics simulation of the liquid- initio theory and experiment for H-chabazite, J. Phys. Chem. C 119 (2015)
metal amorphous-semiconductor transition in germanium, Phys. Rev. B 49 6128–6137.
(1994) 14251–14269. [62] B.A. De Moor, M.F. Reyniers, G.B. Marin, Physisorption and chemisorption of
[33] G. Kresse, J. Furthmuller, Efficient iterative schemes for ab initio total-energy alkanes and alkenes in H-FAU: a combined ab initio-statistical
calculations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169– thermodynamics study, Phys. Chem. Chem. Phys. 11 (2009) 2939–2958.
11186. [63] C.J. Cramer, Essentials of Computational Chemistry: Theories and Models,
[34] G. Kresse, J. Furthmuller, Efficiency of ab-initio total energy calculations for second ed., Wiley, Chichester Hoboken, NJ, 2004.
metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 [64] A.C. Hindmarsh, ODEPACK, a systematized collection of ODE solvers, in: R.S.
(1996) 15–50. Stepleman et al. (Eds.), IMACS Transactions on Scientific Computation, vol. 1,
[35] P.E. Blochl, Projector augmented-wave method, Phys. Rev. B 50 (1994) 17953– North-Holland, Amsterdam, 1983, pp. 55–64.
17979. [65] S.L. Boyd, R.J. Boyd, P.W. Bessonette, D.I. Kerdraon, N.T. Aucoin, A theoretical-
[36] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector study of the effects of protonation and deprotonation on bond-dissociation
augmented-wave method, Phys. Rev. B 59 (1999) 1758–1775. energies, J. Am. Chem. Soc. 117 (1995) 8816–8822.
M. John et al. / Journal of Catalysis 330 (2015) 28–45 45

[66] S.R. Blaszkowski, R.A. vanSanten, The mechanism of dimethyl ether formation [72] M.T. Aronson, R.J. Gorte, W.E. Farneth, An infrared-spectroscopy study of
from methanol catalyzed by zeolitic protons, J. Am. Chem. Soc. 118 (1996) simple alcohols adsorbed on H-ZSM-5, J. Catal. 105 (1987) 455–468.
5152–5153. [73] D. Ramey, S.-T. Yang, Production of Butyric Acid and Butanol from Biomass,
[67] A.J. Jones, E. Iglesia, Kinetic, spectroscopic, and theoretical assessment of Final Report to the US Department of Energy, 2004.
associative and dissociative methanol dehydration routes in zeolites, Angew. [74] M. D’Amore, L. Manzer, E. Miller, R. DiCosimo, J. Knapp, Process for Making
Chem. Int. Ed. 53 (2014) 12177–12181. Butenes from Aqueous 1-Butanol, U.S. Patent 20080015395, 2008.
[68] H. Chiang, A. Bhan, Catalytic consequences of hydroxyl group location on the [75] J.F. DeWilde, H. Chiang, D.A. Hickman, C.R. Ho, A. Bhan, Kinetics and
rate and mechanism of parallel dehydration reactions of ethanol over acidic mechanism of ethanol dehydration on gamma-Al2O3: the critical role of
zeolites, J. Catal. 271 (2010) 251–261. dimer inhibition, ACS Catal. 3 (2013) 798–807.
[69] V. Van Speybroeck, J. Van der Mynsbrugge, M. Vandichel, K. Hemelsoet, D. [76] C.B. Phillips, R. Datta, Production of ethylene from hydrous ethanol on H-ZSM-
Lesthaeghe, A. Ghysels, G.B. Marin, M. Waroquier, First principle kinetic 5 under mild conditions, Ind. Eng. Chem. Res. 36 (1997) 4466–4475.
studies of zeolite-catalyzed methylation reactions, J. Am. Chem. Soc. 133 [77] W.R. Moser, R.W. Thompson, C.C. Chiang, H. Tong, Silicon-rich H-ZSM-5
(2011) 888–899. catalyzed conversion of aqueous ethanol to ethylene, J. Catal. 117 (1989) 19–
[70] M.F. Reyniers, G.B. Marin, Experimental and theoretical methods in kinetic 32.
studies of heterogeneously catalyzed reactions, Annu. Rev. Chem. Biomol. 5 [78] A. Corma, J. Perezpariente, Catalytic activity of modified silicates: 1.
(2014) 563–594. Dehydration of ethanol catalyzed by acidic sepiolite, Clay Miner. 22 (1987)
[71] M.T. Aronson, R.J. Gorte, W.E. Farneth, The influence of oxonium ion and 423–433.
carbenium ion stabilities on the alcohol/H-ZSM-5 interaction, J. Catal. 98 [79] M.T. Aronson, R.J. Gorte, W.E. Farneth, D. White, Stability of adsorption
(1986) 434–443. complexes formed by alcohols on H-ZSM-5, Langmuir 4 (1988) 702–706.

You might also like