Single-Event Kinetic Model for Cracking and Isomerization of 1

Nov 19, 2014 - This might cause sterically demanding olefins to react before leaving the zeolite pore system. Internal diffusion limitations in the po...
0 downloads 0 Views 1MB Size
Subscriber access provided by University of British Columbia Library

Article

Single-Event Kinetic Model for Cracking and Isomerization of 1-Hexene on ZSM-5 Tassilo von Aretin, and Olaf Hinrichsen Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/ie503628p • Publication Date (Web): 19 Nov 2014 Downloaded from http://pubs.acs.org on November 24, 2014

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.

Industrial & Engineering Chemistry Research 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 39

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

Industrial & Engineering Chemistry Research

Single-Event Kinetic Model for Cracking and Isomerization of 1-Hexene on ZSM-5 Tassilo von Aretin and Olaf Hinrichsen* Technische Universität München, Department of Chemistry, Lichtenbergstraße 4, 85748 Garching b. München, Germany Technische Universität München, Catalysis Research Center, Ernst-Otto-Fischer-Straße 1, 85748 Garching b. München, Germany Single events, kinetic model, olefins cracking, ZSM-5

The single-event kinetic modeling approach is applied to olefins cracking. Experimental data from literature is used for the model development and the identification of reaction pathways in olefins cracking. A detailed product spectrum from a 1-hexene feed on ZSM-5 gives insight into the reactivity through the fractions of 20 different product olefins and the combined C7= - C12= olefins. The reaction network comprising all possible elementary reactions on ZSM-5 is generated using a matrix notation. Single-event rate constants for cracking and isomerization of 1-hexene are determined from parameter estimation. The results show the capability of the modeling approach to reproduce the complex product distribution observed on ZSM-5.

ACS Paragon Plus Environment

1

Industrial & Engineering Chemistry Research

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 2 of 39

Introduction: Both ethene and propene offer a wide range of applications for the production of synthetic materials. These hydrocarbons are mainly produced by steam cracking of naphtha or catalytic cracking of crude oil into petroleum fractions1. Since the demand for propene is growing faster than for ethene, the production of low olefins on demand offers economic advantages2. In this context the catalytic cracking of higher olefins on acid zeolites allows a process design with high propene yield. To obtain further mechanistic insight into the reactions occurring during catalytic cracking of olefins, a kinetic model is developed. Due to the large number of reaction possibilities, the single-event kinetic modeling approach is used to link catalyst design and product distribution in catalytic cracking. The aim of this work is to reproduce the product distribution of olefins cracking with consideration of the complex elementary reactions on ZSM5.

Experimental Data: Since only few kinetic models for olefins cracking are found in literature3,4, experimental data sets suited for the development of a kinetic model are rare. Vahteristo et al.3 study skeletal isomerization of 1-pentene on ZSM-22. Here, however, only the pentene isomers are given in detail with dimerization and fragmentation products regarded as two combined responses3. Borges et al.4 determine rates of consumption for light olefins on ZSM-5, but product distributions are only given at certain temperatures with carbon numbers being lumped. Experimental data for cracking of n-butene solely shows the fractions of main products on ZSM5 with fractions of pentenes or hexenes being combined5,6. Thus, detailed information about the occurring pathways leading to the corresponding isomers is not included in the data set. Also not

ACS Paragon Plus Environment

2

Page 3 of 39

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

Industrial & Engineering Chemistry Research

suited for kinetic modeling are data points which are given over time on stream instead of conversion or residence time7,8. Abbot and Wojciechowski9 investigate catalyst deactivation on ZSM-5 with the experiments being conducted in an integral plug flow reactor9,10. Here, a detailed product spectrum obtained from a 1-hexene feed at 350°C is given depending on conversion in the absence of catalyst deactivation9,11,12. It contains the weight fractions of 20 olefin isomers in the range of C2= to C6= and the combined weight fraction of C7= - C12= olefins9. Thus, this product spectrum allows mechanistic insight into the reaction pathways on ZSM-5 and the extent of occurring dimerization reactions. Three catalyst to reactant mass ratios are used9. The integral conversion  is determined from the conversion at the reactor outlet  13.    =   

(1)



Here is the total duration for which a certain mass of feed is passed through a certain mass of catalyst11. Hence, the residence time in the reactor is dependent on and the catalyst to reactant mass ratio 14.  = 

(2)

Abbot and Wojciechowski9,15 give the integral conversion of the feed 1-hexene depending on . Thus, every data point for one catalyst to reactant mass ratio corresponds to a different residence time in the reactor9,11. Hereby, the catalyst is fully deactivated for large values of since a further increase in residence time does not result in a higher integral conversion9,14. Consequently, a dependence between residence time and conversion, which can be used for the steady state modeling of a plug flow reactor, is not available. The detailed product spectrum, however, can be used for the determination of kinetic parameters for cracking and isomerization of 1-hexene. Absolute values of the rate constants are consequently not covered by the

ACS Paragon Plus Environment

3

Industrial & Engineering Chemistry Research

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 4 of 39

experimental data. The ratios of the rate constants, however, represent the reaction pathways occurring on ZSM-5 and correspond to the experimentally obtained product distribution. A silicium to aluminum ratio of 80 is used by Abbot and Wojciechowski9. From this value the number of acid sites is calculated using the stoichiometric formula NanAlnSi96-nO192 ∙ 16 H2O for ZSM-516 and a stoichiometric coefficient n of 1.2 for Al. Assuming that all aluminum atoms are accessible, the number of acid sites  is calculated to be 0.1976





. Since the product

spectrum is only given at one temperature, it is not possible to determine activation energies9. Instead rate constants for cracking and isomerization are determined by kinetic modeling to show the applicability of the single-event kinetic approach to olefins cracking. According to the experimental data by Abbot and Wojciechowski9, doublebond-isomerization is reported to be a comparatively fast reaction. Thus, conversion of the 1-hexene feed is calculated from any component other than linear hexenes and the weight fractions of the doublebond-isomers of 1-hexene are not listed9,15. For linear pentenes, 3-methylpentenes and partly for 2-methylbutenes as well as for 2-methylpentenes the mole fractions of the doublebondisomers can be plotted over conversion of n-hexene. As shown in figure 1, the mole fractions of the doublebond-isomers are independent of the conversion, indicating equilibrium among these. For the 3-methylpentenes, however, the assumption of equilibrium among the doublebond-isomers is only valid at conversions larger than approximately 20 %. This might be due to 2-ethylbut-1-ene being an impurity in the feed as described by Abbot and Wojciechowski9. Despite of this, equilibrium among doublebondisomers is assumed in the model, which is valid for most olefins over a wide range of conversions. Herby, the olefins shown in figure 1 are regarded to be representative of the product

ACS Paragon Plus Environment

4

Page 5 of 39

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

Industrial & Engineering Chemistry Research

spectrum in olefins cracking. Moreover, an equilibration between doublebond-isomers also indicates a protonation equilibrium.

Figure 1: Mole fractions of various doublebond-isomers of C5= and C6= olefins depending on the conversion of n-hexene9

Reaction pathways for hexene cracking: Abbot and Wojciechowski9,15 identify the unimolecular cracking of C6= to propene and the dimerization of C6= to a C12= intermediate as reaction pathways at the given temperature of

ACS Paragon Plus Environment

5

Industrial & Engineering Chemistry Research

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 6 of 39

350°C. This C12= intermediate subsequently cracks into either C9= and C3=, or C8= and C4=, or C7= and C5=. The olefins produced by cracking can thereby undergo consecutive reactions, like the cracking of C9= to C6= and C3= olefins9,15. By analysis of the liquid residue of all conducted experiments on ZSM-5, Abbot and Wojciechowski9,17 conclude that no olefins larger than C12= are obtained with a hexene feed. Using a pentene feed, C10= is the largest detected hydrocarbon15. Thus, in the reaction network oligomerisation of the feed is regarded to be limited to the formation of dimers. In the liquid residue of the experiments of Abbot and Wojciechowski9 mostly linear and monobranched olefins were detected, with dibranched olefins having the methyl side groups at different carbon atoms. Tabak et al.18 also report a low degree of branching for product olefins in the range of C11= to C20= on ZSM-5, with most branches being methyl side groups. Weitkamp19 investigates the cracking of paraffins on ZSM-5 and concludes that tribranched carbenium ions are not formed due to steric constraints. Abbot and Wojciechowski9 do not list olefins with quaternary carbon atoms in their product spectrum on ZSM-5. Hereby, even the dibranched olefins 2,3-dimethlybut-1-ene and 2,3-dimethlybut-2-ene are observed only in small amounts, which is regarded to be due to pore size restrictions9. Thus, species with quaternary carbon atoms are excluded from the reaction network generation due to steric constraints on ZSM-5. Branching is limited to a maximum of two methyl side groups and the maximum carbon number considered in the elementary reactions is twelve. Furthermore a low weight fraction of C7= - C12= olefins is detected in the experimental product spectrum9. The thermodynamic data at 350°C, however, allows the formation of large olefins in considerable amounts. Thus, in the model a free gas-phase equilibrium is not assumed. Instead dimerization pathways via secondary educt and tertiary product carbenium ions are excluded

ACS Paragon Plus Environment

6

Page 7 of 39

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

Industrial & Engineering Chemistry Research

from the reaction network due to steric constraints in the pores of ZSM-5. Only dimerization pathways via secondary carbenium ions are included in the model and cracking via tertiary to secondary carbenium ions is regarded as an irreversible reaction. Further dimerization reactions of product olefins like e.g. propene, butenes and pentenes with each other and the feed are also considered thus accounting for the complex reaction possibilities between olefins and carbenium ions on the acid zeolite. According to Abbot and Wojciechowski9 ethene is only formed as a secondary product at high conversions of hexene. Buchanan20 studies the cracking of hexene at 538°C on ZSM-5 and identifies propene and butene as the main products. Ethene is regarded to be formed from pentene at long residence times20. Buchanan et al.21 investigate hexene cracking on ZSM-5 at 510°C and determine reaction rates for the different cracking modes. Hereby, the rate of ethene formation is determined to be low compared to propene formation due to the low stability of the primary carbenium ions involved21. When considering cracking of an octene feed on ZSM-521, faster cracking modes via the more stable tertiary and secondary carbenium ions are available. Thus, the slower cracking pathways to ethene become insignificant21. In the model ethene is regarded to be only formed from the cracking of linear pentenes. It is considered to be a slow reaction compared to cracking via secondary or tertiary carbenium ions. For linear pentenes, however, cracking to ethene and propene is the only available cracking pathway. Assuming cracking to ethene being a slow reaction corresponds to the experimental observation that ethene is only formed in low amounts at high conversion9.

Reaction network generation:

ACS Paragon Plus Environment

7

Industrial & Engineering Chemistry Research

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 8 of 39

To incorporate the reactivity of olefins into a kinetic model, a reaction network including all possible elementary reactions occurring on the acid catalyst is generated. A matrix notation assigning a unique Boolean relation matrix to every olefin and carbenium ion considered is applied22,23,24,25. Three-dimensional matrices are used, with the first plane describing the carbon skeletal structure, the second the doublebond structure and the third the charge structure of the molecule. This is depicted in table 1 for the but-2-yl carbenium ion and isobutene.

Table 1: Matrix notation for but-2-yl carbenium ion and isobutene

Carbon skeletal

Doublebond

structure

structure

C1 C2 C3 C4

Charge structure

C1 C2 C3 C4

C1 C2 C3 C4

But-2-yl

C1

0

1

0

0

C1

0

0

0

0

C1

0

0

0

0

carbenium

C2

1

0

1

0

C2

0

0

0

0

C2

0

1

0

0

ion

C3

0

1

0

1

C3

0

0

0

0

C3

0

0

0

0

C4

0

0

1

0

C4

0

0

0

0

C4

0

0

0

0

C1 C2 C3 C4

Isobutene

C1 C2 C3 C4

C1 C2 C3 C4

C1

0

1

0

0

C1

0

1

0

0

C1

0

0

0

0

C2

1

0

1

1

C2

1

0

0

0

C2

0

0

0

0

C3

0

1

0

0

C3

0

0

0

0

C3

0

0

0

0

C4

0

1

0

0

C4

0

0

0

0

C4

0

0

0

0

ACS Paragon Plus Environment

8

Page 9 of 39

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

Industrial & Engineering Chemistry Research

The first row of the carbon skeletal structure of the but-2-yl carbenium ion in table 1 indicates a bond between carbon atom C1 and carbon atom C2. Since no further C-C bonds are present on carbon atom C1, the rest of this row contains zeroes. The second row shows a bond between carbon atom C2 and C1 as well as C3. Doublebonds are defined analogously in the doublebond structure. Positive charges only correspond to one carbon atom, as indicated for carbon atom C2 in the charge structure of the but-2-yl carbenium ion. Forming the sum over the second row in the carbon skeletal structure of isobutene in table 1 gives three, thus indicating the tertiary carbon atom in isobutene. Elementary reactions are interpreted as matrix operations by deleting or creating bonds in the matrices. In the reaction network every product has also been considered as a reactant, thus ensuring that all reaction possibilities are taken into account. The matrix notation allows the computer aided generation of the reaction network independent of the considered olefin. The reactivity of olefins on ZSM-5, on which the reaction network generation is based, is shown in figure 2 for exemplary olefins26,27,28. Due to the equilibrium between doublebond isomers observed in figure 1, the hydride shift reactions in figure 2 (b) are also equilibrated. For PCP-branching reactions shown in figure 2 (c) both alpha and beta PCP-branching are considered27.

ACS Paragon Plus Environment

9

Industrial & Engineering Chemistry Research

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 39

Figure 2: Elementary reactions with transition states considered in the reaction network generation for olefins on ZSM-5: (a) protonation/deprotonation, (b) hydride shift, (c) methyl shift, (d) PCP-branching with protonated cyclopropane (PCP) ring as transition state and (e) cracking and dimerization

To handle the large number of parameters necessary for the kinetic description of the reaction network, the single-event kinetic model is applied23,25,28,29. Using the single-event approach the rate constant for every elementary reaction is divided into two parts30,31,32. The first part accounts for the different types of carbenium ions involved, e.g. secondary or tertiary ions. This procedure subdivides the elementary reactions into different reaction families, which are described by only one rate constant depending on the reactant and product carbenium ion25,33,34. Hereby, this part of

ACS Paragon Plus Environment

10

Page 11 of 39

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

Industrial & Engineering Chemistry Research

the rate constant consists of an enthalpic as well as an intrinsic entropic contribution and is called the single-event rate constant35. The second part of the rate constant considers the entropy difference between reactant and transition state by accounting for the structure of molecules via global symmetry numbers23,30,32. Structural differences between elementary reactions, like e.g. the position of methyl side groups in a molecule, are comprehended by this second part of the rate constant called the number of single events35. Global symmetry numbers are calculated from the matrix notation described above by determining external and internal rotational axes in the molecule as well as the number of chiral centers23,36,37. With the conformations of transition states taken from literature, which are shown in figure 2, the number of single events is obtained from these global symmetry numbers for every elementary reaction in the network27,28,38. Thus, multiple elementary reactions are comprehended by a limited amount of kinetic parameters, since only the single-event rate constants have to be determined from an experimental data set.

Rate Equations: The concept of rate determining steps is used to describe the reaction pathways in the acidcatalyzed cracking of olefins. Cracking, dimerization, methyl shift and PCP-branching reactions are considered as rate determining steps due to equilibria observed in the experimental data. The reaction rate of a carbenium ion of the type m, i.e. tertiary, secondary or primary, into a product carbenium ion of type n is given below for the considered rate determining steps.  ; ! =

$

" # ;

! &

(3)

%

$

'(/** ; ! =

" #'(/** ;

+,' ; ! =

$

" #+,' ;

! & %

! & -./ %

(4) (5)

ACS Paragon Plus Environment

11

Industrial & Engineering Chemistry Research

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 12 of 39

For isomerization and cracking the rate equations are implemented as first-order reactions in the concentration of carbenium ions on the surface34,39. Dimerization is considered to be the reaction between a carbenium ion adsorbed on the zeolite and a gas-phase olefin. Thus, the rate of this rate determining step depends on the concentration of the carbenium ion and the partial pressure of the educt olefin. From a protonation equilibrium the concentration of the carbenium ion & is determined depending on the concentration of the physisorbed olefin .% . %

01%

& = 0 %

2& %

456 76 ; ! 3 489 78 ; 76 ! . 3 %

(6)

According to the single-event kinetic approach a reference olefin is applied for the protonation equilibrium23,25. As shown by Vynckier and Froment25 this procedure is necessary to avoid thermodynamic inconsistencies in the description of equilibria between doublebond-isomers. For pentenes and larger olefins the 2-methylalkenes are the reference olefins, whereas for smaller olefins the 1-alkenes are used28,34. The physisorption of gas-phase olefins is implemented via a Langmuir approach39.  :;,1% 51%

.% =  @ ∑

(7)

B :;,1B 51B

Thus, the reaction rate of every olefin can be calculated depending on the gas-phase partial pressures. Hereby, a reaction pathway proceeds via equilibrated physisorption and protonation steps to the rate determining step. Desorption reactions following the rate determining step, leading to the product olefins are also equilibrated34. The rate of formation of an olefin C.% is obtained by summing up the reaction rates of the different reaction pathways. E+,' C.% = ∑E DF D  ; G , 78 ! − ∑DF D  , 78 ; G ! + C & %

(8)

By the sum over the number of cracking reactions C the formation of olefin 78 from the arbitrary carbenium ions  and G is considered. The consumption of 78 is taken into account

ACS Paragon Plus Environment

12

Page 13 of 39

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

Industrial & Engineering Chemistry Research

via the sum over all dimerization reactions JKL. Reaction rates of carbenium ions stemming from olefin 78 are comprehended by C& , which is explained below: %

C& = ∑ '(/**  ; 8 ! − ∑ '(/** 8 ;  ! + ∑   ; 8 , 7M ! − %

∑  8 ;  , 7M ! + ∑ +,'  , 7M ; 8 ! − ∑ +,' 8 , 7M ;  !

(9)

Hereby, all carbenium ions in the reaction network are considered by the summation over the index k. Rate determining steps forming the carbenium ion i are added to the rate of formation of carbenium ion C8@ , others consuming it are subtracted. The olefin 7M is arbitrary and does not affect the rate of the rate determining step for cracking reactions34. To obtain the flow rates of every olefin in the reaction network, the design equation for an isothermal plug flow reactor is solved34. NO1% NP

= C.%

(10)

This ordinary differential equation is solved by integrating the molar flow Q.% of the olefin 78 over the mass of catalyst R with the ode15s solver in MATLAB.

Equilibria in the derivation of the reaction rate: Cracking and dimerization can be considered as forward and reverse reactions. Thus, the rate constants for cracking and dimerization are not independent from each other. Instead the rate constants for two pathways involving the same olefins have to follow the constraint that after long residence times the gas-phase equilibrium is established.

ACS Paragon Plus Environment

13

Industrial & Engineering Chemistry Research

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 39

Figure 3: Reaction pathways for cracking and dimerization splitted into forward and reverse reactions to determine an equilibrium constant for the rate determining step

Figure 3 shows one dimerization pathway starting from the gas-phase olefins 78 and 7M over the rate determining step, leading to olefin 7S . The reaction pathway for the backward reaction starts from olefin 7S . To express the rate constant for dimerization via the rate constant for cracking, an equilibrium constant for the rate determining step is used. 3+( =

TUV E;!

(11)

W2 ;E!

In figure 3 the equilibrium constant for the rate determining step is obtained by multiplication of the equilibrium constants for physisorption, protonation and between the gas-phase olefins. Thus, the educts of the rate determining step are linked with the product carbenium ion via the gas-phase equilibrium constant. 3+( F :



;XY .% ;E!

:



;,1%

3"G 35Z[9,.\ 356 7S ; !

(12)

ACS Paragon Plus Environment

14

Page 15 of 39

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

Industrial & Engineering Chemistry Research

Here, the protonation equilibrium also has to be described with the corresponding reference olefins. To calculate the single-event rate constant for dimerization from the single-event rate constant for cracking, all equilibrium constants have to be single-event equilibrium constants. This can be derived by equating the reaction rates for cracking and dimerization, thus yielding the equilibrium condition. Hence the following equation results: 4

4

4+( F :]^ :;,1\% :;XY 3 4 4 `. ,Ea : : ;XY

X

4%>Y_ .\ ;.X\ ! .X\ ,! :



% %>Y_ `.% ;.X a :;,1%

(13)

Here, all symmetry numbers cancel out when converting equilibrium constants to single-event equilibrium constants23. The different reference olefins are labeled, being 768 the reference olefin corresponding to olefin 78 . Since no change in the symmetry numbers is assumed for physisorption, the equilibrium constant for physisorption is assumed to be equal to the corresponding single-event equilibrium constant. The advantage of this procedure is a reduction in the number of kinetic constants to be determined from parameter estimation. The same equilibrium considerations which describe cracking and dimerization as forward and backward reaction have to be applied for PCP-branching32. Here, elementary reactions via the single-event rate constants #$**  ; b! and #$** b; ! are also forward and backward reactions32. Thus, these two single-event rate constants are linked by a thermodynamic constraint32. Derived from a scheme similar to figure 3, the single-event rate constant for #$**  ; b! is expressed via #$** b; !, the equilibrium constants in both reaction pathways and the gas phase equilibrium constant between the educt and product olefins of the isomerization pathway. 4

$

>

4

>

4

. . ! > :;XY X ;9! :%>Y_ > ;.X 4+( = cWc ;9! = :]^ :;,1 3 $ 4 4 `.  ;a : `. ;. a :  9;! : cWc

;XY

X

%>Y_



X

;,1

(14)

Hereby, symbols from the forward reaction pathway proceeding from olefin 7 via a tertiary educt carbenium ion are labeled with , whereas symbols of the backward reaction pathway with

ACS Paragon Plus Environment

15

Industrial & Engineering Chemistry Research

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 16 of 39

olefin 79 as educt over a secondary carbenium ion are labeled with b. Equating the reaction rates of both pathways for PCP-branching with #$**  ; b! and #$** b; ! yields the correct equilibrium condition between the olefins. The number of differential equations is reduced by considering equilibrium among doublebond-isomers. Therefore, the molar flows of all doublebond-isomers corresponding to one carbon skeletal structure are combined. At one experimental condition the mole fractions of these doublebond-isomers are determined. To obtain the molar flow of one specific olefin for calculating the reaction rate, the molar flow of its carbon skeletal structure is multiplied with its mole fraction in doublebond-isomerization equilibrium. Thus, e.g. the molar flow of 1-pentene is obtained from the molar flow of linear pentenes multiplied with the mole fraction of 1-pentene in equilibrium among 1-pentene, cis-2-pentene and trans-2-pentene. Consequently, the number of differential equations necessary to calculate the rate of all linear pentenes is reduced to one. When considering larger olefins, like dodecenes, this reduction becomes more advantageous. This procedure is comparable to the a posteriori lumping applied by Cochegrue et al.40 or by Guillaume et al.41, who also consider equilibria between isomerization reactions. In the calculation of equilibria, cis- and trans-isomers have to be distinguished. As e.g. for doublebond-isomers like 1-alkenes and 2-alkenes, the enthalpy of cis- and trans-isomers also differs42,43. Thus, an isomer group of cis- and trans-isomers is formed to consider the difference in thermodynamic data44. Δe,f = − Cg h ijk- l

mnf%> o

p + jk- l

mnfXq> o

pr

(15)

This Gibbs' energy of an isomer group Δe,f is taken to determine equilibrium constants in the model when cis- and trans-isomers are involved.

ACS Paragon Plus Environment

16

Page 17 of 39

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

Industrial & Engineering Chemistry Research

Thermodynamic data: For the calculation of equilibrium constants, Gibbs' energies are used. Experimental data from Alberty and Gehrig42 are available for olefins up to hexenes. For larger olefins estimates of the Gibbs' energy are obtained with Benson's group contribution method43,45. To be used together, Gibbs' energies from both sources are referred to the elements hydrogen in the gas-phase and crystalline graphite42,46,47. Thermodynamic data for physisorption and protonation are taken from theoretical calculations by Nguyen et al.48, who consider linear olefins with carbon numbers ranging from two to eight. Linear correlations are given to determine the thermodynamic data for physisorption and chemisorption on ZSM-548. A Si/Al ratio of 95 is investigated by Nguyen et al.48 and thus the data is comparable to the Si/Al ratio of 80 used by Abbot and Wojciechowski9. Differences in the Si/Al ratio are regarded to influence the activity, but for hexene cracking similar product distributions are obtained from different Si/Al ratios on ZSM-5 as shown by Buchanan20. Hereby, Nguyen et al.48 distinguish between 1-alkenes as well as 2/3/4-alkenes and state different linear correlations for these. For protonation this discrimination is applied in the model, but to describe physisorption only the data for 1-alkenes is used. The physisorption data from Nguyen et al.48 corresponds to the one for linear olefins, whereas in the reaction network linear and branched olefins are considered. Here, using the discrimination between 1-alkenes and 2/3/4alkenes for physisorption as proposed by Nguyen et al.48 results in an inferior description of the experimental product spectrum by the kinetic model. Thus, the physisorption parameters determined by Nguyen et al.48 for 1-alkenes seem to be better suited for the description of linear and branched olefins. Nguyen et al.48 consider the protonation of a physisorbed olefin to an alkoxide. Whether the intermediate on zeolites is an alkoxide or a carbenium ion is not finally

ACS Paragon Plus Environment

17

Industrial & Engineering Chemistry Research

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 18 of 39

clarified in literature49,50,51. Regarding the reactivity, however, the same conclusions are obtained for carbenium ions and alkoxides by the single-event approach39. Reaction pathways via activation of tertiary carbon atoms are faster than pathways via secondary ones19,34,39. Since according to Nguyen et al.48 the protonation of a physisorbed alkene leads to an alkoxide, alkoxides are also considered as the reaction intermediates in the presented model. Kazansky et al.52 conclude that alkoxides are the stable intermediates on zeolites, but carbenium ions are regarded as the high energy transition states. Hereby, the stability of primary, secondary and tertiary alkoxides is found to be comparable53 and consequently no stability difference between these has to be considered in the presented model39.

Parameter estimation: The assumptions described above reduce the number of kinetic parameters to be determined from experimental data. With quaternary carbon atoms being excluded from the reaction network, cracking reactions with tertiary product carbenium ions become impossible. The same holds for methyl shift reactions involving tertiary carbenium ions as well as for PCP-branching reactions between tertiary educt and product carbenium ions. A Levenberg-Marquardt algorithm is applied to minimize the difference between molar flows predicted by the model and the experimental data54. Therefore, the solver lsqnonlin is used in MATLAB with ode15s being applied to determine the molar flows of olefins in the reactor model. Boundary conditions on the reactor inlet are set according to the feed composition given by Abbot and Wojciechowski9. Thus, the feed consists of 98.39 % 1-hexene and 1.61 % 2-ethylbut-1-ene9. In total the kinetic model is based on 144 methyl shift, 784 PCP-branching, 141 cracking and 119 dimerization reactions with 593 different olefins being considered. Due to the assumption of equilibrium in

ACS Paragon Plus Environment

18

Page 19 of 39

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

Industrial & Engineering Chemistry Research

doublebond-isomerization, 86 ordinary differential equations have to be solved for the carbon skeletal structures of the olefins in the reaction network. The mass balance of the model and the correct implementation of the equilibrium condition between reaction pathways have been checked. The single-event rate constants are determined from the experimental data of Abbot and Wojciechowski9. The ratios of the estimated single-event rate coefficients, which are normalized with #$ b; -!, are given in table 2.

Table 2: Ratios of single-event rate constants estimated from the experimental data according to Abbot and Wojciechowski9 #$ b; -! #$ b; -!

#$ b; b! #$ b; -!

#$  ; b! #$ b; -!

#$'( b; b! #$ b; -!

#$** b; b! #$ b; -!

#$**  ; b! #$ b; -!

1.0

2.5 ∙ 10m

1.2 ∙ 10x

8.5 ∙ 10x

1.7 ∙ 10m

5.0

ACS Paragon Plus Environment

19

Industrial & Engineering Chemistry Research

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 20 of 39

With the estimated single-event rate constants from table 2 the product distribution depending on the conversion of n-hexene is calculated and compared to the experimental data from Abbot and Wojciechowski9. This is shown in figure 4 for the C2= to C5= olefins and in figure 5 for the C6= as well as the C7= - C12= olefins

Figure 4: Comparison between experimental product distribution according to Abbot and Wojciechowski9 and the single-event kinetic model with the rate constants from table 2 for the C2= to C5= olefins

ACS Paragon Plus Environment

20

Page 21 of 39

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

Industrial & Engineering Chemistry Research

Figure 5: Comparison between experimental product distribution according to Abbot and Wojciechowski9 and the single-event kinetic model with the rate constants from table 2 for the C6= and the C7= - C12= olefins

The experimental mass fractions of the C2= to C5= olefins in figure 4 are in good agreement with the model predictions. Propene shows a systematic deviation of approximately 0.5 wt% and a slight underprediction is observed for the linear pentenes. Hereby, the equilibrium mole fractions in doublebond-isomerization lead to a slight overprediction of cis-pent-2-en in

ACS Paragon Plus Environment

21

Industrial & Engineering Chemistry Research

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 22 of 39

combination with an underprediction of trans-pent-2-ene. Similar observations are obtained in figure 5 for some isomers of the methylpentenes. According to the experimental data the predicted weight fractions of 2,3-dimethylbutenes show the behavior of a stable secondary product9, but at high conversions the model overpredicts the weight fractions by roughly 0.5 wt%. At medium conversions, an overprediction is also observed for the weight fraction of the C7= - C12= olefins in figure 5.

Discussion of the results: Cracking and isomerization of 1-hexene on ZSM-5 are described according to the single-event kinetic approach. The influence of the catalyst on the reactivity is included into the model via steric constraints. Thus, compared to the free gas-phase chemistry reaction pathways are excluded from the reaction network based on experimental observations. Molecules with a high degree of branching or quaternary carbon atoms are excluded from the reaction network since these are not experimentally detected by Abbot and Wojciechowski9,15. This affects the parameter estimates as having more reaction pathways with the same rate constants would result in different formation rates for the separate olefins. On the contrary it is also possible that these molecules are formed as intermediates in the zeolite pores, but react before they diffuse out of the catalyst. Haag et al.55 determine the diffusion coefficients for linear hexenes, 3methylpentenes

and

2,2-dimethylbutenes

to

3 ∙ 10m| }~ b m,

4 ∙ 10mx }~ b m,

7∙

10m€ }~ b m, respectively. Consequently, it might also be possible that sterically demanding olefins with e.g. a quaternary carbon atom or more than two methyl side groups are present as reaction intermediates, but have a higher residence time in the zeolite pores due to the lower diffusion coefficient. This might cause sterically demanding olefins to react before leaving the

ACS Paragon Plus Environment

22

Page 23 of 39

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

Industrial & Engineering Chemistry Research

zeolite pore system. Internal diffusion limitations in the pores of ZSM-5 have been investigated by Vandegehuchte et al.56 for n-hexane conversion with a model accounting for intracrystalline diffusion limitations and thus the diffusion coefficients of different hydrocarbons. Here, it is concluded that intracrystalline diffusion limitations but not transition-state shape selectivity or physisorption selectivity are the reason for the observed selectivity effects56. From the experimental data on olefins cracking available it does not seem possible to conclude whether highly branched olefins or olefins containing quaternary carbon atoms exist as intermediates or not. Thus, these are excluded from the reaction network in order to only consider reaction pathways leading to experimentally detected olefins. Only dimerization pathways via secondary carbenium ions are included in the reaction network. Consequently, cracking of tertiary educt to secondary product carbenium ions is an irreversible reaction. This results from the conducted parameter estimations. Otherwise a significantly larger amount of C7= - C12= olefins compared to the experimental data is obtained by the kinetic model. At 350°C, the gas-phase equilibrium favors the formation of large olefins. Abbot and Wojciechowski9, however, detect only a small amount of C7= - C12= olefins in their product spectrum. This supports the assumption that an equilibrium composition of olefins as in the free gas-phase is not obtained. Instead, it is assumed that these reaction pathways are blocked due to steric constraints and thus excluded from the reaction network. Hereby, a maximum is observed for the modeled C7= - C12= olefins in figure 5, which is not clearly identifiable from the data of Abbot and Wojciechowski9. Since the C7= - C12= olefins are regarded as reaction intermediates which undergo subsequent cracking, a maximum in their weight fraction is reasonable. Judged from the weight fraction of C7= - C12= olefins, the single-event kinetic model comprehends the extent of dimerization reactions on ZSM-5.

ACS Paragon Plus Environment

23

Industrial & Engineering Chemistry Research

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 24 of 39

Considering the ratios of the single-event rate constants in table 2, increasing values for the single-event rate constants are expected with more stable carbenium ions taking part in the reaction28,34,57. This is the case for PCP-branching and also observed for the two single-event rate constants #$ b; b! and #$  ; b!. The single-event rate constant #$ b; b!, however, is smaller than #$ b; -!. Moreover, the systematic deviation in the weight fraction of propene (cf. figure 4) also indicates that #$ b; b! is too small. Propene is found to be the only olefin in the reaction network which can be formed via #$ b; b! from hexenes without a dimerization step and subsequent cracking21. Thus, a value for #$ b; b!, which is too small, has to result in a lower predicted weight fraction of propene compared to the experimental data in figure 4. Taking into account the relatively small systematic deviation observed for propene of 0.5 wt%, the estimated value for #$ b; b! does not seem to be significantly too low. Thereby, #$ b; b! is linked to the gas-phase equilibrium constant. A higher value for this rate constant directly leads to faster dimerization and a larger weight fraction of C7= - C12= olefins in the product spectrum. This might indicate that too many reaction pathways for dimerization are considered in the reaction network. A further reduction of these pathways, however, can hardly be justified from the experimental data available. Another explanation for the apparently illogical ordering of the estimated single-event rate constants #$ b; -! and #$ b; b! in table 2 might be intracrystalline diffusion limitations in ZSM-5 as investigated by Vandegehuchte et al.56. Taking this into account, it cannot be excluded that the estimated kinetic parameters might also be influenced by diffusion effects in the zeolite pores. Further experimental data seems, however, necessary to conclude on whether intracrystalline diffusion limitations, steric constraints imposed on certain reaction pathways as

ACS Paragon Plus Environment

24

Page 25 of 39

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

Industrial & Engineering Chemistry Research

presented in this paper or combinations of these effects lead to the observed product distribution of 1-hexene on ZSM-5. Dimerization is assumed to proceed via an Eley-Rideal mechanism over a carbenium ion being chemisorbed on the zeolite reacting with a gas-phase olefin. This allows the description of dimerization as the reverse reaction of cracking. A Langmuir-Hinshelwood mechanism with both educts of dimerization adsorbed on the zeolite might also be possible. Data for adsorption are, however, only available for physisorption of an olefin on an empty acid site, which are determined by Nguyen et al.48. In a Langmuir-Hinshelwood mechanism the physisorption of the olefin 7M has to be considered on an acid site on which a hydrocarbon is already adsorbed. Thermodynamic data for this process is not available, since data taken from Nguyen et al.48 is assumed to be different from physisorption data on an acid site covered with an adsorbed hydrocarbon. Hence, the assumption of an Eley-Rideal mechanism allows the use of the data provided by Nguyen et al.48 to describe all sorption processes. As an assumption the physisorption parameters determined by Nguyen et al.48 for 1-alkenes are used to describe the physisorption of all linear and branched olefins in the reaction network. Hereby, the discrimination between 1-alkenes and 2/3/4-alkenes proposed by Nguyen et al.48 is not applied. Using this discrimination, it is observed from the parameter estimation results that the ratios of 2-methylpentenes and 3-methylpentenes would be predicted contrary to the experimental data. Thus, the correlation given by Nguyen et al.48 for 1-alkenes seems to be better suited to describe the physisorption of linear and branched olefins on ZSM-5. The experimental data from Abbot and Wojciechowski9 allows no evaluation of the dependency between residence time and conversion of n-hexene. Consequently, absolute values of the single-event rate constants are not covered by the experimental data from Abbot and

ACS Paragon Plus Environment

25

Industrial & Engineering Chemistry Research

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 26 of 39

Wojciechowski9. The ratios of these rate constants, however, allow to draw conclusions on the overall rate of the different reaction mechanisms. These ratios reflect the reactivity of olefins on ZSM-5, leading to the product spectrum depending on the conversion. Minor deviations in the equilibrium mole fractions in the figures 4 and 5 result from the thermodynamic data used. This leads e.g. to different ratios of cis- and trans-2-pentene in figure 4, but despite of that the total amount of linear pentenes is reproduced by the kinetic model. Given the predicted fractions of linear butenes and methylbutenes in figure 4, the assumption of equilibrium in doublebond-isomerization can be regarded as valid. This also holds for the 2methylpentenes in figure 5. Here, deviations in the weight fractions of cis- and trans-4methylpent-2-ene are also due to the thermodynamic data used since the single-event approach does not discriminate between cis- and trans-isomers23. A comparably large deviation between model and experiment is observed for trans-4-methylpent-2-ene. This, however, is due to the large fraction of the total 2-methylpentenes in the product spectrum. The weight fraction of trans-4-methylpent-2-ene is calculated from the total molar flow of 2-methylpentenes and the mole fraction of trans-4-methylpent-2-ene in doublebond-isomerization equilibrium. Thus, small discrepancies in this equilibrium mole fraction result in comparably large absolute deviations for trans-4-methylpent-2-ene (cf. figure 5). As shown in figure 1, the assumption of equilibrium between doublebond-isomers is not valid for the 3-methylpentenes at conversions below 20 %. This explains the deviations in figure 5 for 2-ethylbut-1-ene, which is an impurity in the feed9. For conversions larger than 20 % the model reproduces the experimentally obtained weight fractions of 3-methylpentenes, showing the suitability of this assumption. Assuming equilibrium in doublebond-isomerization allows the use of the concept of the rate determining step in the

ACS Paragon Plus Environment

26

Page 27 of 39

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

Industrial & Engineering Chemistry Research

derivation of the rate equations as presented. The alternative of estimating kinetic parameters for protonation is not regarded to be possible from the experimental dataset available. Based on experimental observations, the formation of ethene is considered to occur from pentenes20,21. Other reaction pathways leading to ethene are excluded from the reaction network and ethene is also not regarded as an educt for dimerization reactions. This assumption is based on the low reactivity of ethene compared to larger olefins58. Considering the small weight fraction of ethene being formed even at high conversions in figure 4, this assumption seems to be justified under the experimental conditions.

Conclusion: The single-event kinetic modeling approach is capable of describing the complexity of the reactions occurring during cracking and isomerization of 1-hexene on ZSM-5. Hereby, 1188 reaction pathways with 593 different olefins are considered in the reaction network according to which the rate equations are set up. In the model formulation, steric constraints are used to represent the influence of the catalyst on the reactivity of olefins. Based on experimental observations on ZSM-5, reaction pathways are excluded from the reaction network as a result from these steric constraints. The advantage of applying the single-event kinetic approach is the small number of parameters necessary to describe the complex reactivity. This results in an increased stability of the parameter estimation routine used for determining single-event rate constants from the experimental data. Insight into the ratios of reaction rates for different reaction pathways leading to the detailed experimental product distribution on ZSM-5 is obtained from the estimated single-event rate constants. To describe the reaction rates of olefins on the zeolite, the concept of

ACS Paragon Plus Environment

27

Industrial & Engineering Chemistry Research

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 28 of 39

rate determining steps and equilibrated elementary reactions in the different reaction pathways can be used. Hereby, the complex product spectrum of hexene cracking on ZSM-5 with 20 different olefin isomers and the combined C7= - C12= olefins is reproduced by the proposed kinetic model.

Corresponding Author [email protected]

Acknowledgments: The authors acknowledge the financial support from Clariant Produkte (Deutschland) GmbH.

Supporting Information The single-event methodology is applied to a reference case from literature34 to show that the chosen modeling approach is used correctly. This information is available free of charge via the Internet at http://pubs.acs.org/.

Abbreviations CR, cracking; DIM, dimerization; MS, methyl shift; PCP, protonated cyclopropane ring; RDS, rate determining step; List of Symbols:

ACS Paragon Plus Environment

28

Page 29 of 39

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

Industrial & Engineering Chemistry Research

 &

Concentration of carbenium ion C8@ on the zeolite surface

h #‚ƒ„

.%

Concentration of physisorbed olefin 78 on the zeolite surface

h #‚ƒ„



Total concentration of acid sites on the zeolite

Q.%

Molar flow of olefin 78 in the reactor

%

Δeƒ89

Gibbs energy of cis-isomer

Δe,f

Gibbs energy of an isomer group consisting of cis- and trans isomer

Δe6„E9 # ; !

Gibbs energy of trans-isomer

Rate constant for cracking reaction between educt carbenium ion of type  to product carbenium ion of type

#$ ; !

Single-event rate constant for cracking reaction between educt carbenium ion of type  to product carbenium ion of

h #‚ƒ„ h b … h … h … h 1 b 1 b

type #$+,' ; !

Single-event rate constant for dimerization reaction between educt carbenium ion of type  to product carbenium ion of

1 †‡ b

type

ACS Paragon Plus Environment

29

Industrial & Engineering Chemistry Research

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

489 78 ; 76 ! 3

Single-event equilibrium constant for isomerization of olefin

Page 30 of 39

[-]

78 to reference olefin 76 35Z[9,.% 456 76 ; ! 3

Equilibrium constant for physisorption of olefin 78

Single-event equilibrium constant for protonation of reference

1 †‡ [-]

olefin 76 to carbenium ion of type  456 `768 , ma 3

Single-event equilibrium constant for protonation of the

[-]

reference olefin 76 corresponding to olefin 78 to carbenium ion of type  3+(

Equilibrium constant of the rate determining step of a reaction pathway

4+( 3

Single-event equilibrium constant of the rate determining step of a reaction pathway

"

C JKL 

-./

1 †‡ 1 †‡

Number of single events

[-]

Number of all cracking reactions in the reaction network

[-]

Number of all dimerization reactions in the reaction network

[-]

Catalyst to reactant mass ratio

Partial pressure of olefin ‰

#‚ƒ„ #‚ ""N †‡

ACS Paragon Plus Environment

30

Page 31 of 39

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

Industrial & Engineering Chemistry Research

+( ; !

Rate of rate determining step from educt carbenium ion of type  to product carbenium ion of type

C

Universal gas constant

h #‚ƒ„ b … h 3

C.%

Rate of formation of olefin 78

h #‚ƒ„ b

C &

Rate of formation of carbenium ion C8@

h #‚ƒ„ b

Secondary carbenium ion

[-]

Š.%

Symmetry number of olefin 78

[-]

Š.%,B

Symmetry number of olefin ‹ stemming from paraffin Œ

[-]

Š&

Symmetry number of carbenium ion C8@

[-]

Š &

Symmetry number of carbenium ion # stemming from

[-]

%

b

%

%,

paraffin Πg

Temperature

3



Tertiary carbenium ion



Duration of an experiment

b



Residence time in the reactor

b

[-]

ACS Paragon Plus Environment

31

Industrial & Engineering Chemistry Research

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 32 of 39

R

Mass of catalyst in the reactor

#‚ƒ„



Conversion of component A at the reactor outlet

[-]



Integral conversion of component A

[-]

ACS Paragon Plus Environment

32

Page 33 of 39

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

Industrial & Engineering Chemistry Research

References: (1) Ren, T.; Patel, M.; Blok, K. Olefins from conventional and heavy feedstocks: Energy use in steam cracking and alternative processes. Energy. 2006, 31, 425. (2) Mokrani, T.; Scurrell, M. Gas Conversion to Liquid Fuels and Chemicals: The Methanol Route - Catalysis and Processes Development. Cat. Rev. - Sci. Eng. 2009, 51, 1. (3) Vahteristo, K.; Sahala, K.-M.; Laari, A.; Solonen, A.; Haario, H. Skeletal isomerization kinetics of 1-pentene over an HZSM-22 catalyst. Chem. Eng. Sci. 2010, 65, 4640. (4) Borges, P.; Ramos Pinto, R.; Lemos, M. A. N. D. A.; Lemos, F.; Védrine, J. C.; Derouane, E. G.; Ramôa Ribeiro, F. Light olefin transformation over ZSM-5 zeolites: A kinetic model for olefin consumption. Appl. Catal., A 2007, 324, 20. (5) Houžviča, J.; Nienhuis, J. G.; Hansildaar, S.; Ponec, V. ZSM-5 - An active, selective and stable catalyst for skeletal isomerization of n-butene. Appl. Catal., A 1997, 165, 443. (6) Rutenbeck, D.; Papp, H.; Freude, D.; Schwieger, W. Investigations on the reaction mechanism of skeletal isomerization of n-butenes to isobutene Part I. Reaction mechanism on HZSM-5 zeolites. Appl. Catal., A 2001, 206, 57. (7) Guisnet, M.; Andy, P.; Boucheffa, Y.; Gnep, N. S.; Travers, C.; Benazzi, E. Selective isomerization of n-butenes into isobutene over aged H-ferrierite catalyst: nature of the active species. Catal. Lett. 1998, 50, 159.

ACS Paragon Plus Environment

33

Industrial & Engineering Chemistry Research

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 34 of 39

(8) Klepel, O.; Loubentsov, A.; Böhlmann, W.; Papp, H. Oligomierzation as an important step and side reaction for skeletal isomerization of linear butenes on H-ZSM-5. Appl. Catal., A 2003, 255, 349. (9) Abbot, J.; Wojciechowski, B. W. Catalytic Cracking and Skeletal Isomerization of n-Hexene on ZSM-5 Zeolite. Can. J. Chem. Eng. 1985, 63, 451. (10) Corma, A.; Wojciechowski, B. W. A Comparison of HY and LaY Cracking Activity in Cumene Dealkylation. J. Catal. 1979, 60, 77. (11) Best, D.; Wojciechowski, B. W. On Identifying the Primary and Secondary Products of the Catalytic Cracking of Cumene. J. Catal. 1977, 47, 11. (12) Ko, A.-N.; Wojciechowski, B. W. Catalytic Isomerization of 1-Hexene on HY Zeolite. Int. J. Chem. Kinet. 1983, 15, 1249. (13) Wojciechowski, B. W. The Kinetic Foundations and the Practical Application of the Time on Stream Theory of Catalyst Decay. Cat. Rev. - Sci. Eng. 1974, 9, 79. (14) Wojciechowski, B. W. A Theoretical Treatment of Catalyst Decay. Can. J. Chem. Eng. 1968, 46, 48. (15) Abbot, J.; Wojciechowski, B. W. The Mechanism of Catalytic Cracking of n-Alkenes on ZSM-5 Zeolite. Can. J. Chem. Eng. 1985, 63, 462. (16) Broach, R. W. Zeolite Types and Structures. In “Zeolites in Industrial Separation and Catalysis” (Kulprathipanja, S.). Wiley-VCH: Weinheim, 2010.

ACS Paragon Plus Environment

34

Page 35 of 39

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

Industrial & Engineering Chemistry Research

(17) Abbot, J.; Corma, A.; Wojciechowski, B. W. The Catalytic Isomerization of 1-Hexene on ZSM-5 Zeolite: The Effects of a Shape-Selective Catalyst. J. Catal. 1985, 92, 398 (18) Tabak, S. A.; Krambeck, F. J.; Garwood, W. E. Conversion of Propylene and Butylene over ZSM-5 Catalyst. AIChE J. 1986, 32, 1526. (19) Weitkamp, J. Catalytic Hydrocracking - Mechanisms and Versatility of the Process. ChemCatChem 2012, 4, 292. (20) Buchanan, J. S. Gasoline selective ZSM-5 FCC additives: Model reactions of C6 – C10 olefins over steamed 55:1 and 450:1 ZSM-5. Appl. Catal., A 1998, 171, 57. (21) Buchanan, J. S.; Santiesteban, J. G.; Haag, W. O. Mechanistic Considerations in AcidCatalyzed Cracking of Olefins. J. Catal. 1996, 158, 279. (22) Baltanas, M. A; Froment, G. F. Computer Generation of Reaction Networks and Calculation of Product Distributions in the Hydroisomerization and Hydrocracking of Paraffins on PtContaining Bifunctional Catalysts. Comput. Chem. Eng. 1985, 9, 71. (23) Baltanas, M. A.; van Raemdonck, K. K.; Froment, G. F.; Mohedas, S. R. Fundamental Kinetic Modeling of Hydroisomerization and Hydrocracking on Noble-Metal-Loaded Faujasites. 1. Rate Parameters for Hydroisomerization. Ind. Eng. Chem. Res. 1989, 28, 899. (24) Clymans, P. J.; Froment, G. F. Computer-generation of reaction paths and rate equations in the thermal cracking of normal and branched paraffins. Comput. Chem. Eng. 1984, 8, 137.

ACS Paragon Plus Environment

35

Industrial & Engineering Chemistry Research

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 36 of 39

(25) Vynckier, E.; Froment, G. F. Kinetic Modeling of Complex Catalytic Processes based upon Elementary Steps. In “Kinetic and Thermodynamic Lumping of Multicomponent Mixtures” (Astarita, G. and Sandler, S. I.). Elsevier: Amsterdam, 1991. (26) Brouwer D. M. Reactions of Alkylcarbenium Ions in Relation to Isomerization and Cracking of Hydrocarbons. In “Chemistry and Chemical Engineering of Catalytic Processes” (Prins, R. and Schuit G. C.). Sijthoff & Noordhoof: Alphen aan den Rijn, 1980. (27) Martinis, J. M.; Froment, G. F. Alkylation on Solid Acids. Part 2. Single-Event Kinetic Modeling. Ind. Eng. Chem. Res. 2006, 45, 954. (28) Svoboda, G. D.; Vynckier, E.; Debrabandere, B.; Froment, G. F. Single-Event Rate Parameters for Paraffin Hydrocracking on a Pt/US-Y-Zeolite. Ind. Eng. Chem. Res. 1995, 34, 3793. (29) Feng, W.; Vynckier, E.; Froment, G. F. Single-Event Kinetics of Catalytic Cracking. Ind. Eng. Chem. Res. 1993, 32, 2997. (30) Alwahabi, S. M.; Froment, G. F. Single-Event Kinetic Modeling of the Methanol-to-Olefins Process on SAPO-34. Ind. Eng. Chem. Res. 2004, 43, 5098. (31) Froment, G. F. Kinetic modeling of acid-catalyzed oil refining processes. Catal. Today 1999, 52, 153. (32) Surla, K.; Guillaume, D.; Verstraete, J. J.; Galtier, P. Kinetic Modeling using the SingleEvent Methodology: Application to the Isomerization of Light Paraffins. Oil Gas Sci. Technol. 2011, 66, 343.

ACS Paragon Plus Environment

36

Page 37 of 39

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

Industrial & Engineering Chemistry Research

(33) Froment, G. F. Single Event Kinetic Modeling of Complex Catalytic Processes. Cat. Rev. Sci. Eng. 2005, 47, 83. (34) Thybaut, J. W.; Marin, G. B.; Baron, G. V.; Jacobs, P. A.; Martens, J. A. Alkene Protonation Enthalpy Determination from Fundamental Kinetic Modeling of Alkane Hydroconversion on Pt/H-(US)Y Zeolite. J. Catal. 2001, 202, 324. (35) Thybaut, J. W.; Marin, G. B. Single-Event MicroKinetics: Catalyst design for complex reaction networks. J. Catal. 2013, 308, 352. (36) Thybaut, J. W.; Marin, G. B. Kinetic Modeling of the Conversion of Complex Hydrocarbon Feedstocks by Acid Catalysts. Chem. Eng. Technol. 2003, 26, 509. (37) Toch, K.; Thybaut, J. W.; Vandegehuchte, B. D.; Narasimhan, C. S. L.; Domokos, L.; Marin, G. B. A Single-Event MicroKinetic model for “ethylbenzene dealkylation/xylene isomerization” on Pt/H-ZSM-5 zeolite catalyst. Appl. Catal., A 2012, 425-426, 130. (38) Park, T. Y.; Froment, G. F. Kinetic Modeling of the Methanol to Olefins Process. 1. Model Formulation. Ind. Eng. Chem. Res. 2001, 40, 4172. (39) Thybaut, J. W.; Narasimhan, C. S. L.; Marin, G. B.; Denayer, J. F.; Baron, G. V.; Jacobs, P. A.; Martens, J. A. Alkylcarbenium ion concentrations in zeolite pores during octane hydrocracking on Pt/H-USY zeolite. Catal. Lett. 2004, 94, 81. (40) Cochegrue, H.; Gauthier, P.; Verstraete, J. J.; Surla, K.; Guillaume, D.; Galtier, P.; Barbier, J. Reduction of Single Event Kinetic Models by Rigorous Relumping: Application to Catalytic Reforming. Oil Gas Sci. Technol. 2011, 66, 367.

ACS Paragon Plus Environment

37

Industrial & Engineering Chemistry Research

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 38 of 39

(41) Guillaume, D.; Valéry, E.; Verstraete, J. J.; Surla, K.; Galtier, P.; Schweich, D. Single Event Kinetic Modelling without Explicit Generation of Large Networks: Application to Hydrocracking of Long Paraffins. Oil Gas Sci. Technol. 2011, 66, 399. (42) Alberty, R. A.; Gehrig, C. A. Standard Chemical Thermodynamic Properties of Alkene Isomer Groups. J. Phys. Chem. Ref. Data 1985, 14, 803. (43) Benson, S. W.; Cruickshank, F. R.; Golden, D. M.; Haugen, G. R.; O'Neil, H. E.; Rodgers, A. S.; Shaw, R.; Walsh, R. Additivity Rules for the Estimation of Thermochemical Properties. Chem. Rev. 1969, 69, 279. (44) Alberty, R. A.; Oppenheim, I. Analytic expressions for the equilibrium distribution of isomer groups in homologous series. J. Chem. Phys. 1986, 84, 917. (45) Benson, S. W. Thermochemical Kinetics; John Wiley & Sons: New York, 1976. (46) Alberty, R. A.; Gehrig, C. A. Standard Chemical Thermodynamic Properties of Alkane Isomer Groups. J. Phys. Chem. Ref. Data 1984, 13, 1173. (47) Domalski, E. S.; Hearing, E. D. Estimation of the Thermodynamic Properties of C-H-N-OS-Halogen Compounds at 298.15 K. J. Phys. Chem. Ref. Data, 1993, 22, 805. (48) Nguyen, C. M.; De Moor, B. A.; Reyniers, M.-F.; Marin, G. B. Physisorption and Chemisorption of Linear Alkenes in Zeolites: A Combined QM-Pot(MP2//B3LYP:GULP) Statistical Thermodynamics Study. J. Phys. Chem. C 2011, 115, 23831. (49) Boronat, M.; Corma, A. Are carbenium and carbonium ions reaction intermediates in zeolite-catalyzed reactions? Appl. Catal., A 2008, 336, 2.

ACS Paragon Plus Environment

38

Page 39 of 39

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

Industrial & Engineering Chemistry Research

(50) Kazansky, V. B. Adsorbed carbocations as transition states in heterogeneous acid catalyzed transformations of hydrocarbons. Catal. Today 1999, 51, 419. (51) Tuma, C.; Sauer, J. Protonated Isobutene in Zeolites: tert-Butyl Cation or Alkoxide? Angew. Chem. 2005, 117, 4847. (52) Kazansky, V. B.; Frash, M. V.; van Santen, R. A. Quantumchemical study of the isobutene cracking on zeolites. Appl. Catal., A 1996, 146, 225. (53) Rigby, A. M.; Kramer, G. J.; van Santen, R. A. Mechanisms of Hydrocarbon Conversion in Zeolites: A Quantum Mechanical Study. J. Catal. 1997, 170, 1. (54) Kapteijn, F.; Berger, R. J.; Moulijn, J. A. Rate Procurement and Kinetic Modelling. In “Handbook of Heterogeneous Catalysis”. Wiley-VCH: Weinheim, 2008. (55) Haag, W. O.; Lago, R. M.; Weisz, P. B. Transport and Reactivity of Hydrocarbon Molecules in a Shape-Selective Zeolite. Farad. Discuss. 1981, 72, 317. (56) Vandegehuchte, B. D.; Thybaut, J. W.; Marin, G. B. Unraveling Diffusion and Other Shape Selectivity Effects in ZSM5 Using n-Hexane Hydroconversion Single-Event Microkinetics. Ind. Eng. Chem. Res. 2014, 53, 15333. (57) Martens, G. G.; Thybaut, J. W.; Marin, G. B. Single-Event Rate Parameters for the Hydrocracking of Cycloalkanes on Pt/US-Y Zeolites. Ind. Eng. Chem. Res. 2001, 40, 1832. (58) Garwood, W. E. Conversion of C2-C10 to Higher Olefins over Synthetic Zeolite ZSM-5. In “Intrazeolite Chemistry” (Stucky G. D. and Dwyer F. G.). American Chemical Society: Washington, D.C., 1983.

ACS Paragon Plus Environment

39