Diffusion Dependence of the Dual-Cycle Mechanism for MTO

Sep 12, 2017 - The “dual-cycle” pathway (i.e., olefins-based cycle and aromatics-based cycle) of methanol-to-olefin (MTO) has been generally accep...
2 downloads 10 Views 2MB Size
Subscriber access provided by University of Sussex Library

Article

Diffusion Dependence of the Dual-Cycle Mechanism for MTO Reaction Inside ZSM-12 and ZSM-22 Zeolites Zhiqiang Liu, Yueying Chu, Xiaomin Tang, Ling Huang, guangchao Li, Xianfeng Yi, and Anmin Zheng J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b07374 • Publication Date (Web): 12 Sep 2017 Downloaded from http://pubs.acs.org on September 15, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1

Diffusion Dependence of the Dual-cycle Mechanism for MTO

2

Reaction Inside ZSM-12 and ZSM-22 Zeolites

3 4

Zhiqiang Liua,b,#, Yueying Chua,#, Xiaomin Tanga,b,#, Ling Huang a,*,

5

Guangchao Lia,b, Xianfeng Yia,b, Anmin Zhenga,*

6 7

a

8

National Center for Magnetic Resonance in Wuhan, Wuhan Institute of Physics and

9

Mathematics, Chinese Academy of Sciences, Wuhan 430071, P. R. China.

State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics,

b

10

University of Chinese Academy of Sciences, Beijing 100049, P. R. China.

11 12

Corresponding authors: Fax: +86 27 87199291.

13

*

14

E-mail addresses: [email protected]; [email protected]

15

[ # ] These authors contributed equally to this work.

16

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

17

Abstract: The “dual-cycle” pathway (i.e., olefins-based cycle and aromatics-based

18

cycle) of methanol-to-olefin (MTO) has been generally accepted as hydrocarbon pool

19

mechanism. Understanding the role of diffusion of reactant, intermediate and product

20

in the MTO process is essential in revealing its reaction mechanism. By using

21

molecular dynamics (MD) simulations for two one-dimensional zeolites (ZSM-12 and

22

ZSM-22) with a channel difference being only 0.3 Å in pore sizes, the diffusion

23

behaviors of some representative species following “dual-cycle” mechanism (e.g.,

24

methanol, polymethylbenzenes and olefins molecules) have been theoretically

25

investigated in this work. It was found that the diffusion coefficients of methanol and

26

olefins along ZSM-12 were ca. 2~3 times faster than that along ZSM-22 at 673 K. In

27

the aromatics-based cycle, the polymethylbenzenes are crucial intermediates during

28

the MTO reaction. 1,2,3,5-tetramethylbenzene is almost imprisoned inside ZSM-12,

29

such slower diffusion of tetramethylbenzene offers more opportunities for the geminal

30

methylation reaction to form MTO activated pentamethylbenzenium cation, which

31

would split into olefins through “paring” or “side-chain” pathways. However, in the

32

ZSM-22 zeolite, since 1,2,4-trimethylbenzene is stacked, the following methylation

33

reaction solely results in the formation of tetramethylbenzene, which is not a MTO

34

activated species in ZSM-22 and more bulky polymethylbenzene further blocks the

35

channel more seriously. When it comes to the olefins-based cycle, olefins can diffuse

36

freely inside these two zeolites with methoxide intermediate bound to the zeolite

37

frameworks, and thus facilitates formation of longer-chain olefin through olefin

38

methylation reaction in these two zeolite catalysts. Combination of the higher reaction

ACS Paragon Plus Environment

Page 2 of 40

Page 3 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

39

activity (from DFT calculation) and the longer contact time (from MD simulation)

40

between the olefin and methoxide, is apparently illustrated the olefins-based cycle

41

does more preferentially occur inside ZSM-22 than ZSM-12. Apparently, the MTO

42

reaction mechanism is strongly determined by the diffusion behaviors of reaction

43

species inside the zeolite confined pores.

44 45

1. Introduction

46

Owing to the high energy density and easy transportability, olefins play an essential

47

part in the chemical supply chain.1 Nevertheless, the increasing demand for energy

48

and growing depletion of crude oil require the search for new routes for the

49

production of olefins from non-oil sources.2-3 Among these alternative reaction routes,

50

the methanol-to-olefins (MTO) conversion on zeolites has drawn considerable

51

attention from both fundamental research and industrial application in the last several

52

decades.4-8 The interest in the use of methanol as a source feedstock has risen since it

53

can be produced from coal, natural gas, biomass and even CO2.2, 9 Currently, many

54

researches focus on studying the MTO reaction mechanism with the aim to enhance

55

catalyst performance and promote product selectivity.10-12

56

It remains highly challenging to unravel the MTO reaction mechanism and its

57

dependence on the zeolite framework structures, which results from the complexity in

58

product distribution and the difficulty in the identification of active intermediates.5

59

13-14

60

for the formation of light olefins during MTO reaction.15-16 Aromatics (e.g.,

An indirect hydrocarbon pool (HP) mechanism is now gaining wider acceptance

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

61

polymethylbenzenes (PMBs)) and olefins represent two kinds of important HP species

62

for MTO conversion. As illustrated in Scheme 1, the different properties of PMBs and

63

olefins result in two distinct catalytic cycles in principle, known as dual-cycle

64

concept.10, 17-18 During the aromatic-based cycle, light olefins split off from alkyl side

65

chains of cyclic compounds.19-21 In the olefin-based cycle, similar methylation and

66

cracking pathway for aliphatic chains are followed.22-23 Extensive experimental and

67

theoretical results indicated that aromatic species are likely to be the dominating

68

hydrocarbon pool species in H-SAPO-34.24 Nevertheless, a series of contradictory

69

proposals have been arisen on which cycle was followed in some specific zeolite

70

catalysts. For example, it’s experimentally demonstrated by Svelle et al. that ethene is

71

exclusively produced from the aromatic-based cycle, while propene and higher olefins

72

are mainly produced via the olefin-based cycle over H-ZSM-5.10, 18, 25 Therefore, it is

73

crucial for understanding the dependence of the HP species on zeolite frameworks and

74

the relative contribution of each cycle to control the MTO catalytic performance and

75

olefin selectivity.

76 77 78

Scheme 1. The reaction mechanisms of the MTO process in Zeolites. In recent years, studies of the MTO mechanism on two unidimensional ZSM-12

ACS Paragon Plus Environment

Page 4 of 40

Page 5 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

79

and ZSM-22 zeolite catalysts have been debated as well.26-32 ZSM-12 and ZSM-22

80

zeolites possess very similar pore dimensions (5.6 Å × 6.0 Å for ZSM-12 and 4.6 Å ×

81

5.7 Å for ZSM-22) (see Figure 1). Song et al. proposed that 0.3 Å difference in the

82

zeolite pores would cause a dramatic difference in their MTO catalytic

83

performances.14 28 It is revealed that ZSM-12 exhibited a high MTO activity at 673 K,

84

while ZSM-22 was inactive under the same reaction conditions. This difference

85

means that the aromatic hydrocarbon-pool mechanism worked on the ZSM-12 zeolite

86

with the 6.0 Å pore size, while the catalytic cycle was almost inhibited inside the

87

ZSM-22 due to its smaller pore. However, it has been observed experimentally by

88

Svelle29, 33-34 and Liu et al.4, 26-27 that the ZSM-22 zeolite exhibited a great catalytic

89

activity for the MTO reaction through olefin methylation cracking cycle. To reveal the

90

influence of the channel difference for these two zeolites on the reaction routes and

91

catalytic performances, systematic DFT calculations have been performed in our

92

previous work.31 The calculated activation barriers and reaction energies

93

demonstrated that the 0.3 Å channel difference between ZSM-12 and ZSM-22 zeolites

94

lead to a dramatic discrepancy in their transition state selectivity associated with the

95

aromatic-based hydrocarbon pool (HCP) mechanism.

96

It is well known that, besides the activation barriers, the diffusion behaviors of

97

reactant, intermediate and product inside confined pores of zeolite are also essential

98

factors to strongly determine the catalytic performances and product selectivity during

99

catalytic processes.13,

100

35-36

The diffusivity can be measured experimentally by the

pulsed field gradient nuclear magnetic resonance (PFG-NMR) and quasi-elastic

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

101

neutron scattering (QENS) methods.37-38 Recently, microimaging by IR microscopy

102

has been successfully applied to get the diffusivities, by monitoring the distribution of

103

the reactant molecules over the catalyst particles in the moderate experimental

104

conditions.39-40 Nevertheless, it is a challenge to in situ obtain the diffusion coefficient

105

in the reaction conditions (e.g., high temperature and high pressure). As a complement

106

to the experimental approach, the molecular dynamics (MD) theoretical simulation is

107

an effective approach to obtain the diffusion coefficient. It's demonstrated that such

108

state-of-the-art MD calculation has been already used to reveal the influences of

109

temperature and loading on the selectivity of specific hydrocarbon in zeolite

110

catalysts.13 Ghysels and coworkers clarified the dependence of temperature,

111

composition, acidity, and topology on the diffusion of ethene inside microporous

112

zeolites (AEI, CHA, AFX) with 8-ring channel based on the MD simulations.41

113

Furthermore, Wang et al. performed MD to provide an insight into the topology effect

114

on the diffusion of ethene and propene in CHA, MFI, BEA and FAU zeolites with 8-,

115

10- and 12-ring channels.42 Recently, Bu et al. studied the diffusion mechanism of

116

xylene isomers in ZSM-5 using theoretical simulations as well.43

117

As mentioned before, MD calculation has been used to provide the diffusion

118

information of the alkane, olefin and aromatic species inside various zeolite channels

119

with different shapes and sizes.13 However, little systematic work has been done so far

120

for the diffusion of reaction species in the MTO reaction, especially for the

121

co-adsorbed condition with methanol (reactant), polymethylbenzenes/methoxide

122

(intermediate), and olefin (product) inside the zeolite catalysts.

ACS Paragon Plus Environment

Page 6 of 40

Page 7 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

123

In this work, MD simulations were carried out to investigate the diffusion

124

behaviors of reaction species in ZSM-12 and ZSM-22 zeolite materials which have

125

received much interest in the MTO process. Base on the diffusion coefficient, contact

126

time and probability density function, we have established the connection between

127

diffusion behaviors and MTO reaction, and then systematically discussed the

128

diffusion-depended dual-cycle mechanism (aromatics-based and olefins-based cycle)

129

for MTO reaction inside these two zeolites.

130 131

2. Model and Computational Details

132

2.1 Zeolite Frameworks

133 134

Figure 1. Pore size and geometry of (a) MTW along the [010] one-dimensional

135

12-ring channel and (b) TON down the [001] one-dimensional 10-ring channel.

136 137

As shown in Figure 1, ZSM-12 (MTW) and ZSM-22 (TON) are two typical

138

one-dimensional zeolites respectively with the pore sizes of 5.6 Å × 6.0 Å and 4.6 Å ×

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

139

5.7 Å. Notably, a 0.3 Å difference in pore size is present for these two zeolite catalysts.

140

In this work, based on the accessibility of adsorbed molecules, the Si1-O2-Al3 and

141

Si1-O1-Al1 sites were chosen as the locations of Brønsted acidic protons, methoxide

142

(-OCH3) and ethoxide (-OC2H5) species for ZSM-12 and ZSM-22, respectively. (see

143

Figure 1). Initial framework structures of MTW and TON were extracted from the

144

International Zeolite Associations (IZA) database.44 1 × 5 × 2 and 2 × 2 × 5 supercells

145

with the corresponding crystal lattice sizes of 25.6× 26.3 × 24.2 Å3 and 28.2 × 35.7 ×

146

26.3 Å3 were selected to represent ZSM-12 and ZSM-22 zeolites, respectively. There

147

were four acidic protons or alkoxide species (i.e., -OCH3 and -OC2H5) separately

148

distribute inside the different channel in ZSM-12 and ZSM-22 zeolites. During the

149

MD simulation, four guest molecules (e.g., methanol) located inside the different

150

channel of these two zeolites were used for the model of infinite dilution. Furthermore,

151

detailed parameters used in the theoretical calculation were summarized in Table S1.

152 153

2.2 Molecular Dynamics Simulation.

154

All MD simulations were performed in the canonical ensemble (NVT) where the

155

number of particles (N), simulation volume (V), and temperature (T) were kept

156

constant. The simulated temperature was held at 673 K and controlled by a

157

Nosé-Hoover thermostat with a coupling time constant of 1 ps. The velocity Verlet

158

algorithm was used throughout to integrate the Newton’s equations of motion. The

159

consistent valence force field (CVFF), which has been proven to be able to accurately

160

predict the adsorption and diffusion of hydrocarbons in various zeolites materials,45-47

ACS Paragon Plus Environment

Page 8 of 40

Page 9 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

161

was adopted in this work. The long-range electrostatic interactions were calculated

162

using the Ewald summation method48 and Lennard-Jones interactions were calculated

163

with a 9.5 Å cutoff radius.

164

Each MD simulation started with an annealing and followed by 6×106 time steps

165

equilibration with 0.5 fs time step. Then, a production run of 4×107 time steps was

166

performed. At least 5 independent MD simulations were carried out for each system

167

for better statistics and thus the total MD simulation time was 100 ns. The trajectories

168

were recorded every 1000 steps to analyze the mean square displacement, diffusion

169

coefficients and contact time.

170 171 172 173

2.3 Diffusion Coefficient The mean square displacement (MSD) of an adsorbed molecule is defined as the following equation (1) 1 MSD(τ ) = Nm

174

Nm

1 ∑i N τ

2

Nt

∑ [r (t i

0

+ τ ) − ri (t0 )]

(1)

t0

175

where N m is the number of adsorbed molecules, Nτ is the number of time origins

176

used in calculating the average, and ri is the coordinate of the center of mass of

177

molecule i .

178

For one-dimensional diffusion, the slope of the MSD as a function of time

179

determines the self-diffusion coefficient, Ds, defined according to the so-called

180

Einstein relation49

181 182

MSD (τ ) = 2 Dsτ + b

(2)

where b is the thermal factor arising from atomic vibrations. The line is fitted in the

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 40

183

range 50-1000 ps using a least-squares fit. The reported MSD curves and

184

corresponding Ds values in section 3 are calculated as the average of five independent

185

MD trajectories.

186 187 188

2.4 Free Energy The free energy profile is a reliable method to explain the diffusion behavior of

189

hydrocarbon molecule passing through the zeolites.41,

49

190

coordinate of the molecule is chosen along the direction of unidimensional channel in

191

each zeolite, starting with zero and ending at point of the length of the unite cell. Then

192

the normalized histogram of trajectory ξ(t) is the probability distribution P(ξ) of the

193

gas with respects to the direction of channel. Here ξ represents the coordinate of the

194

center of mass of molecule. By taking the logarithm of P(ξ), the free energy profile

195

F(ξ) = -kBTlnP(ξ) is obtained up to an arbitrary constant, where T and kB are the

196

temperature and the Boltzmann constant, respectively.

Firstly, the reaction

197 198

2.5 Contact Time and Probability Density Function

199

For most chemical reactions inside zeolite confined pores, diffusion plays a crucial

200

role to bring the reactants in close proximity, which is also an essential prerequisite

201

for MTO reaction.13 Contact time is defined as the average residence time for the

202

diffused reactant around each active site in certain region, and this parameter can

203

reflect the spatial proximity during the diffusion and reaction.

204

In this work, the contact time between hydrocarbon and active site of zeolites was

ACS Paragon Plus Environment

Page 11 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

205

used to investigate the methylation and dimerization in the MTO reaction. Taking

206

methylation of ethene as an example, the contact time is calculated as following.

207

In order to calculate the contact time, the sizes of zeolites were amplified by N

208

times (i.e., 1×N×1, 1×1×N for ZSM-12 and ZSM-22 respectively) to completely

209

cover the trajectory based on MD simulation of adsorbed molecules. The total time (T)

210

along the trajectory direction was count when the distance between ethene and

211

methoxide is shorter than the certain distance. Therefore, the average contact time

212

(CT) can be calculated by using T/N.

213

The probability density function is defined as

ρ AB ( r

214

)

=

∆Nr + ∆r NA

(3)

215 216

where r is the distance between the active site A and adsorbed molecules B, and NA is

217

the number of active sites A. ∆Nr + ∆r

218

nearest-neighbor of A and located in the distance between r and r + ∆r .

219

means the average over all trajectory.

describes number of B which is the

220 221

Results and Discussion

222

3.1 Diffusion Characteristics of Reactant and Product Inside ZSM-12 and

223

ZSM-22 Zeolites in MTO Reaction

224

As a complement to the experimental approach, the molecular dynamics (MD)

225

simulation is an effective approach to obtain the diffusion coefficient. Since there are

226

lots of reaction species involved in the MTO reaction, such as reactant (methanol),

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

227

products (ethene, propene, butene), intermediate species (p-xylene, trimethylbenzene,

228

tetramethylbenzene, methoxide) and zeolite frameworks (ZSM-12, ZSM-22), it is

229

better to use a generalized force field in the MD calculations. The consistent valence

230

force field (CVFF), which is a generalized valence force field developed by

231

Dauber-Osguthorpe, has been widely used in studying the adsorption and diffusion of

232

gas inside zeolites.45-47 First of all, the reliability of CVFF for the zeolite framework

233

and hydrocarbon species involving the MTO reactions has also been verified (see

234

Figure S1 and Table S2, S3, S4, S5 in Supporting Information).

235

It’s experimentally illustrated that 0.3 Å difference in the pore sizes between

236

ZSM-12 and ZSM-22 zeolite catalysts can lead to the significant difference in the

237

MTO catalytic performances.28 As is well known, the diffusion is an important factor

238

in heterogeneous catalytic processes, especially for the catalytic reaction confined

239

inside the porous catalysts.13, 35 Besides the thermodynamic and kinetic factors as

240

mentioned in our previous work by the DFT calculation, the diffusion of MTO

241

reaction species inside two nano-size zeolite pores strongly mediate the reaction

242

pathways and their reactivities. In order to comprehensively understand the

243

relationship between the zeolite structure and the catalytic performance, it’s necessary

244

to deeply and systematically investigate the diffusion behaviors of reactant (e.g.,

245

methanol), possible olefin products (e.g., ethene, propene and butane) and

246

intermediate (methoxide and PMB), particularly under their co-adsorbed conditions

247

inside these two zeolite catalysts in MTO reaction processes.

ACS Paragon Plus Environment

Page 12 of 40

Page 13 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

248 249

Figure 2. Initial atomic structure and solvent surface of ZSM-12 and ZSM-22 zeolites.

250

Ethene adsorbed in (a) ZSM-12 (along [0 1 0] direction) and (b) ZSM-22 (along [0 0

251

1] direction). Both ethene and p-xylene molecules co-adsorbed in (c) ZSM-12 (along

252

[1 0 0] direction) and (d) ZSM-22 (along [1 0 0] direction).

253 254

The initial structures of ethene adsorbed inside ZSM-12 and ZSM-22 are shown in

255

Figure 2a and 2b. In the MD simulations, five different initial configures were

256

generated after being annealed, then the analyses of mean square displacement (MSD)

257

and diffusion coefficient were carried out by averaging over all five MD results. The

258

MSD of infinitely diluted methanol, ethene, propene, and butene molecules in

259

ZSM-12 and ZSM-22 zeolites under the MTO reaction temperature (673 K) are

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 40

260

displayed in Figure 3. From the slope of MSD, the relative diffusion rate can be

261

qualitatively determined and lead to conclusion that the diffusion of both methanol

262

and olefins in ZSM-12 are faster than that in ZSM-22 zeolite.

263

264 265

Figure 3. MSD with error bars of methanol, ethene, propene and butene molecules in

266

(a) ZSM-12 and (b) ZSM-22 zeolites at 673 K and infinite dilution.

267 268

Table 1. Self-diffusion coefficient (Ds) with standard error in parentheses and isosteric

269

heat (Qst) of hydrocarbon in ZSM-12 and ZSM-22 zeolites at 673 K and infinite

270

dilution. Ds (10-8· m2·s-1) Molecule methanol ethene propene 1-butene

ZSM-12 11.15 10.72 8.10 5.19

(1.17) (1.48) (0.88) (1.09)

Qst (kcal·mol-1)

ZSM-22 4.65 3.29 3.47 1.47

ZSM-12 (1.07) (0.23) (0.45) (0.42)

7.3 9.8 12.7 15.9

ZSM-22 7.9 10.9 14.0 17.5

271 272

Furthermore, the self-diffusion coefficient (Ds) for gas inside zeolites can be

273

extracted using the Einstein relation by equation (2). Thus, on the basis of MSD

ACS Paragon Plus Environment

Page 15 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

274

predicted by MD calculations, the according Ds for methanol, ethene, propene, and

275

1-butene molecules in ZSM-12 and ZSM-22 zeolites can be determined quantitatively.

276

As listed in Table 1, Ds for methanol in ZSM-12 and ZSM-22 are 11.15×10-8 and

277

4.65×10-8 m2/s, respectively. Apparently, 0.3 Å difference in the pore size of ZSM-12

278

zeolite has led to ca. 2.5 times enhancement in the methanol diffusion in comparison

279

with that inside ZSM-22 zeolite. Similar trends are also observed for the olefins

280

(ethene, propene and butene) whose diffusion coefficients range from 10.72×10-8 to

281

5.19×10-8 m2/s in ZSM-12 while from 3.29×10-8 to 1.47×10-8 m2/s inside ZSM-22. It

282

is generally accepted that the pore size plays a vital role in the diffusivity in the

283

heterogeneous catalysis, and the gas diffuse faster in micropores zeolites with

284

relatively larger pores.50 Moreover, the kinetic diameters of adsorbed olefins with the

285

similar structure are important parameter to mediate the diffusion in the zeolites.13 It’s

286

generally accepted that a molecule with the smaller kinetic diameters and molecular

287

weight always leads to relatively faster diffusion in the same zeolite catalyst. Taking

288

olefin as an example, the diffusion coefficients for ethene and propene inside the

289

ZSM-12 zeolite were 10.72×10-8 and 8.1×10-8 m2/s, respectively, and this tendency is

290

in inverse proportion to their kinetic diameters (ethene: 3.9 Å < propene: 4.5 Å).51 It

291

should be pointed out that although butene (4.5 Å) has a similar kinetic diameter with

292

propene, the relatively larger molecular weight leads to its relatively slower diffusion

293

(5.19×10-8 m2/s) inside ZSM-12 zeolite.51 Interestingly, the diffusion of propene

294

(3.47×10-8 m2/s) is found to be a little higher than that of ethene (3.29×10-8 m2/s) in

295

ZSM-22 zeolite, which might be attributed to the resonant diffusion. Resonant

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

296

diffusion, as a unique property for gas molecule diffusion inside confined space (such

297

as zeolites) when the length of molecule matches the lattice result in low activation

298

energy barrier for diffusion, has been studied by both experiment and theory.52 It’s

299

well known that, there are low- and high-energy areas inside the confined space of

300

zeolites. For small gas molecule with short chain, such as methanol, it locates at either

301

low-energy area or high-energy area in the channel, and the diffusion barrier can be

302

obtained by the energy difference of these two areas. During the diffusion process,

303

methanol must overcome this barrier. However, when the end-to-end chain length of a

304

molecule matches the lattice periodicity (in the case of resonant diffusion), the

305

molecule synchronously occupied both low- and high-energy areas of the zeolite,

306

hence would experience a relatively lower diffusion barrier. Exactly, propene rather

307

than ethane molecule meets the requirement of the resonant diffusion, therefore, it’s

308

accepted that a relatively faster diffusion of propene compared with ethene inside

309

ZSM-22 zeolites. Yoo et al. provided experimental evidence for resonant diffusion of

310

normal alkanes in ZSM-12 and Schuring et al. also found the enhanced diffusivity of

311

n-octane in ZSM-22 at 333 K by MD simulation.52-54

312

313

ACS Paragon Plus Environment

Page 16 of 40

Page 17 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

314

Figure 4. (a) Scatter diagram of the diffusion coefficients and isosteric heat of

315

methanol, ethene, propene and butane molecules. (b) Free energy of methanol and

316

propene in ZSM-12 and ZSM-22 zeolites.

317 318

On the other hand, the diffusion behavior can be understood from the interactions

319

between host (zeolite framework) and guest (adsorbed molecule) systems as well. It’s

320

well known that isosteric heat is directly connected with the sorption affinity of guest

321

molecules in zeolites at the Henry regime, where the guest-host interactions govern

322

the sorption. As shown in Table 1, the isosteric heat are in the order: methanol