Catalytic Consequences of Spatial Constraints and Acid Site Location for Monomolecular Alkane Activation on Zeolites

Similar documents
Consequences of Confinement in Zeolite Acid Catalysis. Rajamani Pachayappan Gounder. A dissertation submitted in partial satisfaction of the

Catalytic Consequences of Spatial Constraints and Acid Site Location for Monomolecular Alkane Activation on Zeolites

Understanding Brønsted-Acid Catalyzed Monomolecular Reactions of Alkanes in Zeolite Pores by Combining Insights from Experiment and Theory

Direct Synthesis of H 2 O 2 on AgPt Octahedra: The Importance of Ag-Pt Coordination for High H 2 O 2 Selectivity

Theta-1 zeolite catalyst for increasing the yield of propene when cracking olefins and its potential integration with an olefin metathesis unit

Characterization of nanopores by standard enthalpy and entropy of adsorption of probe molecules

Consequences of Surface Oxophilicity of Ni, Ni-Co, and Co Clusters on Methane. Activation

SUPPORTING INFORMATION. Framework and Extraframework Tin Sites in Zeolite Beta React Glucose Differently

The Roles of Entropy and Enthalpy in Stabilizing Ion-Pairs at Transition States in Zeolite Acid Catalysis

Effects of Zeolite Pore and Cage Topology on Thermodynamics of n-alkane Adsorption at Brønsted Protons in Zeolites at High Temperature

Supporting Information. First principles kinetic study on the effect of zeolite framework on 1-butanol dehydration

= k 2 [CH 3 *][CH 3 CHO] (1.1)

Error in the Estimation of Effective Diffusion Coefficients from Sorption Measurements*

SUPPORTING INFORMATION. and Mark E. Davis*

CHAPTER 2. Structure and Reactivity: Acids and Bases, Polar and Nonpolar Molecules

Supplementary Information. The role of copper particle size in low pressure methanol synthesis via CO 2 hydrogenation over Cu/ZnO catalysts

University of Bucharest, Faculty of Chemistry, Regina Elisabeta Blvd. 4-12, Bucharest, Romania

Probe Reactions of Alcohols and Alkanes for Understanding Catalytic Properties of Microporous Materials and Alumina Oxide Solid Acid Catalysts

BAE 820 Physical Principles of Environmental Systems

DFT Investigation of Alkoxide Formation from Olefins in H-ZSM-5

Supporting Information

Chain Length Effects of Linear Alkanes in Zeolite Ferrierite. 2. Molecular Simulations 1

Part B: Unraveling the mechanism of catalytic reactions through kinetics and thermodynamics

Supplementary information

Journal of Catalysis

NMR and X-ray Diffraction Studies of Phases in the Destabilized LiH-Si System

ADSORPTION IN MICROPOROUS MATERIALS: ANALYTICAL EQUATIONS FOR TYPE I ISOTHERMS AT HIGH PRESSURE

Kinetic Characterisation of Zeolite Catalysts Using Cracking, Alkylation and Other Chemical Reactions

Supplementary Material

Permeation of Hexane Isomers across ZSM-5 Zeolite Membranes

An Introduction to Chemical Kinetics

Isotopic Tracer Studies of Propane Reactions on H-ZSM5 Zeolite

CHAPTER 7. for two different Diels-Alder-dehydration reactions catalyzed by the Lewis acid

Specificity of sites within eight-membered ring zeolite channels for the carbonylation of methyls to acetyls

Complexity in Complex Mixtures For each process chemistry there will be only O(10):

Supporting Information

Definitions and Concepts

HONOUR SCHOOL OF NATURAL SCIENCE. Final Examination GENERAL PHYSICAL CHEMISTRY I. Answer FIVE out of nine questions

In situ 1 Hand 13 C MAS NMR study of the mechanism of H/D exchange for deuterated propane adsorbed on H-ZSM-5

Chemical Kinetics and Reaction Engineering

BAE 820 Physical Principles of Environmental Systems

ANSWER KEY. Chemistry 25 (Spring term 2016) Midterm Examination

7a. Structure Elucidation: IR and 13 C-NMR Spectroscopies (text , , 12.10)

Topic 4 Thermodynamics

4) Interpret in words the equation: P4O10 (s) + 6 H2O (l) 4 H3PO4 (aq)

IUCrJ (2014). 1, , doi: /s

10/26/2010. An Example of a Polar Reaction: Addition of H 2 O to Ethylene. to Ethylene

Thermal Diffusion and Partial Molar Enthalpy Variations of n-butane in Silicalite-1

The reaction whose rate constant we are to find is the forward reaction in the following equilibrium. NH + 4 (aq) + OH (aq) K b.

Structure and Preparation of Alkenes: Elimination Reactions

Chemistry 111 Syllabus

N. Hansen,*, R. Krishna, J. M. van Baten, A. T. Bell,*, and F. J. Keil

Transition State Enthalpy and Entropy Effects on Reactivity. and Selectivity in Hydrogenolysis of n-alkanes

Calculate a rate given a species concentration change.

Oxygen Reduction Reaction

Efficient, scalable and solvent-free mechanochemical synthesis of the OLED material Alq 3 (q = 8-hydroxyquinolinate) Supporting Information

4. NMR spectra. Interpreting NMR spectra. Low-resolution NMR spectra. There are two kinds: Low-resolution NMR spectra. High-resolution NMR spectra

Genesis of Brønsted Acid Sites during Dehydration of 2-Butanol on Tungsten Oxide Catalysts

ALKANE CRACKING IN ZEOLITES: AN OVERVIEW OF RECENT MODELING RESULTS

Electronic Supplementary Information

Diffusion of propylene adsorbed in Na-Y and Na-ZSM5 zeolites: Neutron scattering and FTIR studies

Dehydrogenation of Propane to Propylene Over Pt-Sn/Al 2 O 3 Catalysts: The influence of operating conditions on product selectivity

CHEM 251 (4 credits): Description

Adsorption Equilibria. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

CHE 611 Advanced Chemical Reaction Engineering

Hydrogen Adsorption and Storage on Porous Materials. School of Chemical Engineering and Advanced Materials. Newcastle University United Kingdom

Rapid, Efficient Phase Pure Synthesis of Ca 2 AlNO 3 Layered Double Hydroxide

Experiment 7: Dynamic NMR spectroscopy (Dated: April 19, 2010)

Quantifying hydrogen uptake by porous materials

The Seeding of Methane Oxidation

SUPPLEMENTARY INFORMATION

Sectional Solutions Key

Silver-Decorated Hafnium Metal-Organic Framework for. Ethylene/Ethane Separation

Journal of Catalysis

Brønsted/Lewis Acid Synergy in Methanol-to-Aromatics Conversion. on Ga-Modified ZSM-5 Zeolites as Studied by Solid-State NMR

Mechanistic Insights into Metal-Lewis Acid Mediated Catalytic Transfer. Hydrogenation of Furfural to 2-Methylfuran

A mass spectrometer can be used to distinguish between samples of butane and propanal. The table shows some precise relative atomic mass values.

Foundations of Chemical Kinetics. Lecture 32: Heterogeneous kinetics: Gases and surfaces

c) fitting of the NMR intensity in dependence of the recycle delays 4

Controlling the Isolation and Pairing of Aluminum in Chabazite Zeolites Using Mixtures of Organic and Inorganic Structure-Directing Agents

UNIVERSITY OF CAMBRIDGE INTERNATIONAL EXAMINATIONS Cambridge International Level 3 Pre-U Certificate Principal Subject

Supporting Information

Spin-spin coupling I Ravinder Reddy

Chemical Reactions and Chemical Reactors

AUTOMOTIVE EXHAUST AFTERTREATMENT

Adsorption Processes. Ali Ahmadpour Chemical Eng. Dept. Ferdowsi University of Mashhad

2. What is the charge of the nucleus in an atom of oxygen-17? (1) 0 (2) 2 (3) +8 (4) +17

Local Structure of Framework Aluminum in Zeolite H-ZSM-5 during Conversion of Methanol Investigated by In Situ NMR Spectroscopy

Supporting Information: Ethene Oligomerization in Ni-containing Zeolites: Theoretical Discrimination of Reaction. Mechanisms

Lecture Notes Chem 51A S. King

Unit 11 Instrumentation. Mass, Infrared and NMR Spectroscopy

Chapter 3. Distinguishing between Reaction Intermediates and. Spectators: A Kinetic Study of Acetone Oxidation Using

Specificity of Sites within Eight-Membered Ring Zeolite Channels for Carbonylation of Methyls to Acetyls

Initial carbon carbon bond formation during synthesis gas conversion to higher alcohols on K Cu Mg 5 CeO x catalysts

Supporting Information. for. Angew. Chem. Int. Ed. Z Wiley-VCH 2003

8. ELECTROCHEMICAL CELLS. n Electrode Reactions and Electrode Potentials a. H 2 2H + + 2e. Cl 2 + 2e 2Cl. H 2 + Cl 2 2H + + 2Cl ; z = 2

Organic Chemistry is the chemistry of compounds containing.

Supporting Online Material (1)

Study Guide for Final Exam, Ch , Chem1B, General Chemistry II

Index to Tables in SI Units

Transcription:

Catalytic Consequences of Spatial Constraints and Acid Site Location for Monomolecular Alkane Activation on Zeolites Rajamani Gounder and Enrique Iglesia Supporting Information S.1. 27 Al MAS NMR Spectroscopy of Zeolite Samples. Relative amounts of aluminum in framework and extra-framework locations were determined from integrated areas of peaks respectively centered at 55 ppm (tetrahedral) and 0 ppm (octahedral), referenced to a 1.0 M aqueous solution of Al(NO 3 ) 3, in their 27 Al MAS NMR spectra (Fig. S.1). NMR spectra were collected at the Caltech Solid State NMR Facility on a ruker Avance 500 MHz spectrometer in a wide bore 11.7 Tesla magnet. 27 Al MAS NMR spectra were measured at 130.35 MHz using a 4 mm CPMAS probe with the application of strong proton decoupling and with a magic angle spinning rate of 13 khz. NMR spectra were recorded at ambient temperature by averaging 512 scans with a 0.5 µs pulse and a 6 s delay. Zeolite samples were packed into a 4mm ZrO 2 rotor and were fully hydrated by placing them in a desiccator containing a 1.0 M KCl aqueous solution (~100% relative humidity) for at least 48 h prior to sealing the rotors with a kel-f cap and acquiring NMR spectra. S.2. Infrared Spectroscopic Studies of Zeolite Samples Upon Exposure to CO at 123 K. Experimental methods for the acquisition of infrared spectra, adapted from previously reported methods [S1], are given in Section 2.2. Infrared spectra taken after successive CO doses to H-MOR-S (Fig. S.2) show that CO initially interacted only with S1

strong (L 1 ; extra-framework Al 3+ ) and weak (L 2 ; penta-coordinated Al 3+ ) Lewis acid sites, producing infrared bands for C-O stretches centered at 2224 cm -1 and 2196 cm -1, respectively [S2]. CO interacted with rønsted acid sites through hydrogen bonding (2173 cm -1 ) [S2] and with zeolite channel walls via van der Waals interactions (2137 cm - 1 ) [S2] only after L 1 and L 2 sites were saturated with CO (Fig. S.2). The density of L 1 and L 2 sites (per g) among different zeolite samples are proportional to their integrated peak areas after normalization to those for Si-O-Si overtones (2100-1750 cm -1 ). Infrared spectra upon saturation of Lewis centers by CO at 123 K for three H-MOR samples of varying provenance are shown in Figure S.3. Turnover rates for C 3 H 8 cracking and dehydrogenation (per g) did not vary systematically with either L 1 or L 2 integrated peak areas among different H-MOR (Fig. S.4) or H-MFI (Fig. S.5) samples. S.3. Assessment of Transport Corruptions of Kinetic Measurements. The Mears criterion was used to assess intraparticle mass transport limitations by assuming isothermal and spherical particles. This criterion asserts that intraparticle concentration gradients are negligible when the inequality: Rvrp 1 C D < s e (S.1) is satisfied at the conditions of the experiments. Here, R v is the observed reaction rate per unit volume, r p is the crystallite radius, C s is the reactant concentration at the crystallite surface, and D e is the effective reactant diffusivity [S3]. The largest measured rate constants were ~0.005 mol (mol H + ) -1 s -1 bar -1 (Table 2) and typical experimental conditions for monomolecular alkane reactions (0.2 bar alkane, 780 K, H-MFI, Si/Al = 15) correspond to a per volume reaction rate (R v ) of ~0.2 mol s -1 S2

m -3. The crystallite radius (r p ) was conservatively assumed to be 1000 nm; scanning electron microscopy (SEM) images (not included) indicated that crystallites were 100-250 nm in their longest dimension. The reactant concentration at the external crystallite surface (C s ) was estimated to be 0.3 mol m -3 at 0.02 bar(c 3 H 8 ) using the adsorption constant at 780 K on H-MFI (K ads ~ 0.005 bar -1 ), which was calculated using Eq. S.20 and reported adsorption parameters for C 3 H 8 on H-MFI ( H ads = -46 kj mol -1, S ads = -102 J mol -1 K -1 ) [S4], the saturation loading of C 3 H 8 on H-MFI (q sat ~ 1.5 mol kg -1 ) [S4], and the framework density of MFI (ρ ~ 1.8 x 10 6 g m -3 ). We note that typical acid site densities (C H+ ~ 1 x 10-3 mol(h + ) g -1 ; H-MFI, Si/Al = 15) and intrazeolite reactant concentrations (C C3H8,z ~ 0.3 mol m -3 ) result in a fractional site coverage of ~0.0002, indicating that acid sites remain predominantly unoccupied at monomolecular reaction conditions (i.e., high temperatures, low alkane partial pressures). The effective reactant diffusivity (D e ) was estimated to be 1 x 10-4 cm 2 s -1, a conservative estimate because D e = 3 x 10-4 cm 2 s -1 for n-c 6 H 12 at 811 K in MFI [S5]. These values result in a ratio of reaction to diffusion rates smaller than 10-4 in Eq. S.1 and satisfy the Mears criterion for neglecting intracrystallite mass transport limitations and the resulting concentration gradients. Corruptions arising from heat transfer limitations are expected to be insignificant because of the differential conditions used and the endothermic nature of monomolecular cracking and dehydrogenation reactions. Kinetic measurements were also independent of catalyst bed volume, indicating that homogeneous reactions do not contribute to measured reaction rates and selectivities. S3

S.4. Transition State Treatments of Monomolecular Reaction Rate Laws and the Derivation of Activity Coefficients for Adsorbed Molecules Described by Langmuir Isotherms. Monomolecular reactions of alkanes on rønsted acid sites occur via the pathways shown in Scheme S.1. Alkanes in the gas phase ( g ) adsorb onto rønsted acid sites (H + ) located within the zeolite channel ( z ) in a quasi-equilibrated step. Transition state treatments define reaction rates by the frequency (ν) at which activated complexes ( ), formed upon reactions of adsorbed alkanes with rønsted acid sites (H + ), cross the activation barrier: r = νc, (S.2) In thermodynamically non-ideal systems, the conversion of reactants into the activated complex is given by an equilibrium constant (K ) that depends on the activities of the relevant species: K a γ C = =. (S.3) a a γ C γ C + + + z z z H Z H Z H Z Combining Eqs. S.2 and S.3 results in the following rate expression: kt G γ + C + γ C H Z H Z z z r = exp h RgT γ, (S.4) which can be recast in terms of enthalpies ( H ) and entropies ( S ) of activation kt γ + γ H Z S H z r = C + exp exp C H Z h γ R g RgT z. (S.5) The concentrations (per active site; H + ) (C i ) of adsorbed species ( z, ) are related to their real (P ) or hypothetical (P ) gas-phase pressures by Langmuir isotherms: S4

C i = KiPi 1 + K P + K P. (S.6) Alkanes adsorbed within zeolite channels ( z ) are in equilibrium with alkanes in the gasphase ( g ) at a given pressure (P ) when their chemical potentials are equal: µ = µ. (S.7) z g Assuming that gaseous alkanes behave ideally, their chemical potentials are defined as o P µ = µ ln g + RT g o P, (S.8) where µ o g is the standard gas-phase chemical potential at the reference pressure (P o ) of 1 bar. Chemical potentials of adsorbed species can be analogously written as a o z µ = µ ln z + RT z o, (S.9) a z where a z is the activity of the adsorbed alkanes and µ o z is their standard chemical potential at the reference activity a o z. The combination of Eqs. S.7-S.9 results in the following relationship o o o P a z µ µ ln z = RT g o, (S.10) P a z that defines the dimensionless equilibrium constant (K eq ), K eq o P a P a z =, (S.11) o z which can be further related to K : a o z = K P a, (S.12) z where K eq = K *P o. S5

Activities are related to concentrations through activity coefficients (γ): a = γ c, (S.13) z z z and combining Eqs. S.6, S.12 and S.13 results in the following expression for γ: ( 1 ) γ = a + K P + K P. (S.14) o z z A similar derivation follows for the activity coefficients of transition states, whose surface concentrations are in equilibrium with hypothetical gas-phase pressures and are related by a respective Langmuir adsorption constant. Activity coefficients for alkanes and transition states in Eq. S.5 (γ z, γ ) are identical and cancel because alkanes and hypothetical transition states in the gas-phase competitively adsorb onto the same site (H + ). Furthermore, monomolecular alkane activation requires temperatures that favor low intrazeolite concentrations (C z ) in which H + sites are predominantly unoccupied and both C H + Z and γ H + Z approach unity. Reaction rates therefore become strictly proportional to intrazeolite alkane concentrations (C z ): kt S H r = exp exp C h R g RgT z. (S.15) These rates can be rewritten in terms of pre-exponential (A) and activation energy ( H ) components: H r = Aexp C RgT z. (S.16) Intrazeolite concentrations (C z ) become proportional to external pressures (P ) and to the adsorption constants (K ) in the low-coverage limit (Eq. S.6) and reaction rates (Eq. S.16) then become: r = kint K P, (S.17) S6

with measured rate constants (k meas ): k = k K, (S.18) meas int and temperature dependences for k int and K : k int ( E int / RT ) = Ainte, (S.19) K e e e ( Gads / RT ) ( Hads / RT ) ( Sads / R) = =, (S.20) where H ads and S ads are the enthalpy and entropy of adsorption, respectively. Measured activation energies and pre-exponential factors are given by: E = E + H, (S.21) meas int ads ln( A ) = ln( A ) + ( S / R), (S.22) meas int ads S = S + S, (S.23) meas int ads where E meas and A meas ( S meas ) are measured rate parameters referenced to gas phase alkanes, and E int and A int ( S int ) are intrinsic rate parameters referenced to adsorbed alkanes. We define the measured entropy of activation as: S = R [ln( A ) ln( k T / h)], (S.24) meas meas where A meas is rigorously normalized by the number of acid sites and the number of bonds available for each reaction. This measured activation entropy, although calculated, is simply a redaction of A meas and therefore a measured quantity. S.5. Least-Squares Estimation of Rate Constants for Monomolecular Propane Activation in 8-MR and 12-MR Locations of MOR. The individual contributions of OH groups within 8-MR and 12-MR in H-MOR to measured rate constants were estimated by weighing their respective rate constants by S7

the fraction of OH sites within each environment (Eq. 11). These rate constants were determined by a least-squares regression of rate data on seven MOR samples with different acid site distributions (Table 1). This system of equations can be expressed in matrix notation: x + x + k H,8 MR, H100Na0MOR Z H,12 MR, H, 100Na0MOR Z meas H100Na0MOR Z x + kmeas, H H,8 MR, H83Na17MOR 83Na17MOR Z x + Z H,12 MR, H83Na17MOR Z k meas, H73Na27MOR Z x + x + H,8 MR, H73Na27MOR Z H,12 MR, H73Na27MOR Z k kmeas,8mr meas, H59Na41MOR Z =,8, 59 41,12, 59 41 x + x + H MR H Na MOR Z H MR H Na MOR Z k meas,12 MR, (S.25) kmeas, H45Na55MOR Z x x + + H,8 MR, H45Na55MOR Z H,12 MR, H45Na55MOR Z kmeas, H MOR T x + x + H,8 MR, H MOR T H,12 MR, H MOR T k meas, H MOR S x + x + H,8 MR, H MOR S H,12MR, H MOR S where kmeas is a column vector (7 x 1) containing measured rate constants, x is a matrix (7 x 2) containing the fraction of rønsted acid sites within 8-MR and 12-MR locations, and k loc is a column vector (2 x 1) of measured first-order rate constants in 8-MR and 12- MR locations. Regressed rate constants are shown in Table S.1 at different temperatures, with uncertainties reported as twice the standard deviation. Dehydrogenation rate constants in 12-MR locations are much smaller (by factors of > 10) than rate constants in 8-MR locations and are zero within the error of the regression; these rates were set to zero and the 8-MR dehydrogenation rate constants were regressed again to give final values (shown in Table S.1). S.6. Estimation of Adsorption Enthalpy and Entropy of Propane and n-utane Within 8-MR Pockets of H-MOR, Assuming Complete Confinement. S8

Calorimetric measurements of alkane adsorption on acidic zeolites showed that both isosteric heats of alkane adsorption and the incremental heat of adsorption per methyl or methylene group correlated strongly with the effective pore radius of the zeolite channel [S6]. Here, we attempt to estimate differences in the adsorption enthalpy of C 3 H 8 and n-c 4 H 10 between the 12-MR channel and the 8-MR pocket of H-MOR, assuming that these molecules can be fully contained within 8-MR side pockets. The effective pore diameter, defined to span the centers of the framework oxygen atoms, was calculated by averaging the minor and major axes of the elliptical channel after adding the oxygen atom diameter (0.27 nm). In MOR, the effective pore radii of the 12-MR channel (0.70 x 0.65 nm) and 8-MR pocket (0.26 x 0.57 nm) are 0.47 nm and 0.34 nm, respectively. We note that the effective pore radius for the 8-MR pocket in MOR (0.34 nm) is very similar to the radii for the 8-MR channel (0.34 nm) and 10-MR channel (0.37 nm) in FER, and make the reasonable assumption that the adsorption enthalpy for CH 4 in the 8-MR pocket of MOR is the same as in H-FER ( H ads = -27.7 kj mol -1 ) [S6]. The incremental adsorption enthalpy per CH 2 group in H-FER ( Q st = 12.8 kj mol -1 ) predicts and adsorption enthalpy of -53.3 kj mol -1 for C 3 H 8 in the 8-MR pocket of H-MOR, in agreement with the value reported for C 3 H 8 in H-FER [S6]. Therefore, adsorption parameters for C 3 H 8 in 8-MR locations of H-MOR will be estimated as those reported for adsorption in H-FER ( H ads = -49 kj mol -1, S ads = -103 J mol -1 K -1 ) [S7]. Adsorption parameters for C 3 H 8 in the 12-MR channel of H-MOR were directly measured by calorimetric, gravimetric and infrared spectroscopic methods ( H ads = -41 kj mol -1, S ads = -85 J mol -1 K -1 ) [S4]. Similarly, if n-c 4 H 10 were to be completely contained within 8-MR side pockets of H-MOR, we predict adsorption enthalpies would S9

resemble values measured in H-FER ( H ads = -59 kj mol -1 ) [S7]. The adsorption enthalpy of n-c 4 H 10 in 12-MR H-MOR channels was directly measured to be -50 kj mol -1 [S4]. S.7. Estimation of Kinetic Diameters of Propane and n-utane and 8-MR Pocket Depth in MOR. The kinetic diameters of propane and n-butane [S8] are shown in Scheme S.2. oth molecules have a minimum kinetic diameter of 0.43 nm, which corresponds to their minimum cross-sectional diameters. Propane and n-butane have minimum equilibrium diameters of 0.65 nm and 0.83 nm, respectively, which correspond to their longest dimension. The MOR structure viewed along [001] is shown in Scheme S.3 [S9]. A rough estimate for the depth of the 8-MR side pocket was taken to be half the distance between the edges of adjacent 12-MR channels (0.37 nm) by scaling against the small diameter of the 12-MR ring (0.65 nm). S.8. Distribution of utene Isomers Observed During n-utane Dehydrogenation on MOR Samples. utene isomer ratios measured in the reactor effluent upon n-butane dehydrogenation on H 100 Na 0 MOR-Z and H 45 Na 55 MOR-Z samples at 748 K are plotted against inverse space velocity in Figure S.6. Product chromatograms were corrected for feed butene impurities when calculating dehydrogenation turnover rates, but were not corrected for these impurities when calculating isomer ratios because they equilibrate with the butene products of n-butane dehydrogenation reactions. S10

n-utane dehydrogenation turnover rates (per total H + ) were lower (by a factor of ~7) on H 45 Na 55 MOR-Z than on H 100 Na 0 MOR-Z (Table 6) because the former sample contained a smaller fraction of rønsted acid sites within 8-MR locations (0.13) than the latter (0.56), which were found to exclusively catalyze dehydrogenation turnovers (Section 3.3). utene isomer distributions at 748 K, however, were similar on H 100 Na 0 MOR-Z and H 45 Na 55 MOR-Z (Fig. S.6), in spite of large differences in dehydrogenation turnover rates (per total H + ) on the two samples The position of initial C-H bond activation cannot be preserved in the alkene products because linear butene isomers equilibrate via rapid protonation-deprotonation with intervening hydride shifts. The ratio of isobutene to 1-butene isomers depends on the space velocity and is not equilibrated upon extrapolation to zero conversion, but it is non-zero only because of isobutene impurities present in the feed. This ratio approaches the equilibrated ratio (3.7 at 748 K) [S10] at longer residence times (Fig. S.6), indicating that skeletal rearrangements are not as rapid as double-bond isomerization events at typical reaction temperatures (700-800 K). References [S1] P. Cheung, A. han, G. J. Sunley, D. J. Law, E. Iglesia, J. Catal. 245 (2007) 110. [S2] L. enco, T. ucko, J. Hafner, H. Toulhoat, J. Phys. Chem. 108 (2004) 13656. [S3] D. E. Mears, Ind. Eng. Chem. Process Des. Develop. 10 (1971) 541. [S4] F. Eder, M. Stockenhuber, J. A. Lercher, J. Phys. Chem. 101 (1997) 5414. [S5] W. O. Haag, R. M. Lago, P.. Weisz, Faraday Discussions (1981) 317. [S6] S. Savitz, F. Siperstein, R. J. Gorte, A. L. Myers, J. Phys. Chem. 102 (1998) 6865. [S7] F. Eder, J. A. Lercher, J. Phys. Chem. 101 (1997) 1273. [S8] D. W. reck, Zeolite Molecular Sieves, Wiley, New York, 1974. p. 633-641. [S9] MOR structure viewed along [001] as shown on the International Zeolite Assocation website: http://www.iza-online.org/ [S10] Calculated using thermodynamic data from: D. R. Stull, E. F. Westrum, G. C. Sinke, The Chemical Thermodynamics of Organic Compounds, Wiley, New York, 1987. S11

Scheme S.1. Reaction sequence for monomolecular reactions of alkanes ( denotes a reactant base and P denotes products) on rønsted acid sites (H + ) located within zeolite channels. 1. g z 2. z + H + Z P Scheme S.2. Kinetic diameters of propane and n-butane. Scheme S.3. MOR crystal structure viewed along [001]. S12

Table S.1. Regressed first-order rate constants for cracking and dehydrogenation of propane in the 8-MR pockets and 12-MR channels of MOR at different temperatures; uncertainties reported as twice the standard error. Corrected 8-MR dehydrogenation rate constants were calculated by forcing the 12-MR rate constants to be zero. Temperature (/ K) k meas (8-MR) (/ 10-3 mol (mol H + ) -1 s -1 bar -1 ) k meas (12-MR) (/ 10-3 mol (mol H + ) -1 s -1 bar -1 ) Cracking Dehyd. Dehyd (corr.) Cracking Dehyd. 718 0.7 ± 0.2 0.8 ± 0.3 0.8 ± 0.2 0.3 ± 0.1 0.05 ± 0.2 733 1.2 ± 0.3 1.5 ± 0.6 1.7 ± 0.3 0.4 ± 0.2 0.1 ± 0.4 741 1.6 ± 0.4 2.2 ± 0.8 2.4 ± 0.5 0.5 ± 0.3 0.2 ± 0.6 748 2.0 ± 0.5 2.9 ± 1.2 3.2 ± 0.7 0.7 ± 0.4 0.3 ± 0.8 756 2.6 ± 0.7 4.0 ± 1.7 4.4 ± 0.9 0.9 ± 0.5 0.4 ± 1.2 763 3.4 ± 0.8 5.4 ± 2.2 5.9 ± 1.3 1.1 ± 0.6 0.5 ± 1.6 778 5.5 ± 1.3 9.7 ± 4.2 10.8 ± 2.4 1.7 ± 0.9 1.0 ± 3.0 S13

Figure S.1. 27 Al MAS NMR spectra of zeolite samples. Corresponding extra-framework aluminum content for each sample is given in Table 1. S14

Figure S.2. Infrared spectra of H-MOR-S upon incremental CO doses at 123 K. S15

Figure S.3. Infrared spectra of (a) H-MOR-T, (b) H-MOR-S and (c) H-MOR-Z upon saturation of L 1 and L 2 sites with CO at 123 K. S16

Figure S.4. Propane cracking (, ) and dehydrogenation (, ) turnover rates (per g) measured at 748 K on H-MOR-T, H-MOR-S and H-MOR-Z samples plotted against infrared band areas (normalized to Si-O-Si overtone band areas; 2100-1750 cm -1 ) for L 1 (closed symbols) and L 2 (open symbols) Lewis acid centers upon saturation by CO at 123 K. S17

Figure S.5. Propane cracking (, ) and dehydrogenation (, ) turnover rates (per g) measured at 748 K on H-MFI-1, H-MFI-2, H-MFI-3, and H-MFI-4 samples plotted against infrared band areas (normalized to Si-O-Si overtone band areas; 2100-1750 cm -1 ) for L 1 (closed symbols) and L 2 (open symbols) Lewis acid centers upon saturation by CO at 123 K. S18

Figure S.6. Ratios of cis-2-butene/trans-2-butene (, ), 2-butene/1-butene (, ) and iso-butene/1-butene (, ) isomers in the reactor effluent at 748 K on H 100 Na 0 MOR-Z (closed symbols) and H 45 Na 55 MOR-Z (open symbols) are plotted against inverse space velocity. S19