Anda di halaman 1dari 2

Mechanism of isopropanol dehydration on -Al2O3:

experiments and DFT for a predictive kinetic model


Kim Larmier,1,2 Cline Chizallet,1,* Andr Nicolle, 3 Sylvie Maury,1 Nicolas Cadran,1 Johnny Abboud,2
Anne-Flicie Lamic-Humblot,2 Eric Marceau,2 Hlne Lauron-Pernot2
1
IFP Energies nouvelles, Catalysis and Separation Division, Rond-Point de lchangeur de Solaize, BP3, 69360 Solaize, France
2
Sorbonne Universits, UPMC Univ Paris 06, UMR CNRS 7197, Laboratoire de Ractivit de Surface, 4 place Jussieu, F-75005
Paris, France
3
IFP Energies nouvelles, Powertrain and Vehicule Division, 1-4 avenue de Bois-Prau, 92852 Rueil-Malmaison Cedex, France
*Corresponding author: celine.chizallet@ifpen.fr

Keywords: -alumina, alcohol dehydration, isopropanol, DFT, kinetics, kinetic modeling

1. Introduction calculations (DFT+D2) were performed using the


Alcohols are important platform molecules obtained VASP software (PBE+D2, PAW, Ecutoff = 400 eV),
from the conversion of lignocellulosic biomass. It is and transition state determinations were performed
well known that dehydration reactions of alcohols using the NEB method. Vibrational analyses were
toward both olefins and ethers are successfully performed to deduce free energy profiles. We
catalyzed by common and cost-effective acidic employed the models for the -alumina surface
oxides such as -alumina. Both types of products, proposed by Digne et al.[5] Reactor modeling was
currently supplied by the petrochemicals industry, performed with the Chemkin software, by
are industrially significant, as monomers for implementing the kinetic parameters of the
polymer production (olefins) or as fuel additives elementary steps calculated ab initio.
(ethers).
However, the key parameters governing each 3. Results and discussion
dehydration reaction, in terms of active sites and Isopropanol conversion was measured at 200C over
mechanisms, are still under debate, as shown by three related aluminas. While the specific activity
several experimental[1,2] or computational[3] varied with the catalyst (-Al2O3 being the most
studies dealing with ethanol dehydration. We active one), the selectivity towards diisopropylether
address here the selectivity issue stemming from the was found to be identical at a given conversion (Fig.
two competing dehydration paths offered to 1, marks), with an increase to a maximum for an
isopropanol on -Al2O3 and related solids, and to isopropanol conversion of 35%, and a subsequent
rationalize the evolution of the olefin-to-ether ratio decrease to the benefit of propene.
over the whole conversion range of the alcohol. To
this end, we have set up a combined experimental
and theoretical approach, in which experimental
measurements are compared with the results of DFT
calculations (modeling at the molecular scale) and
microkinetic simulations (modeling at the
macroscopic scale).[4]

2. Experimental Part and Theoretical Details


Experimental measurements were carried out on a
gas-phase catalytic test using a plug-flow reactor.
Before reaction, aluminas (-alumina, Puralox Figure 1. Evolution of partial pressures with an
TH100/150, Sasol ; Na-poisoned - alumina; - equivalent of contact time on -alumina, experimental
alumina) were activated at 450 C (5C.min-1) under (dots) versus simulated (lines) with our final model
nitrogen flow (20 mL.min-1) for 3 h. Isopropanol (reaction temperature 200 C, initial P(iPOH) = 1.5kPa,
(Sigma-Aldrich, 99%) was introduced with a partial mcat = 30 mg).
pressure of 1.5 kPa in a N2 flow (5 to 60 cc.min-1).
The composition of the effluent was determined by This leads to the conclusion that the reaction paths
gas chromatography (Carbowax column). Ab initio giving propene and diisopropylether should not be
treated as parallel and independent reactions in
modeling, and that the active sites should be similar
regardless of alumina, though more or less abundant.
In order to identify plausible active sites, the
reaction pathways leading to the formation of
propene or diisopropylether from isopropanol were
calculated by DFT on the most abundant surfaces of
-alumina, taken in their relevant hydration state
under reaction conditions: (100) facets are
dehydrated while (110) terminations are partially
hydrated (around 9.0 OH.nm-2). The most favorable
pathways for both reactions have been found on the
(100) surface, with an identical intermediate, an
isopropanol molecule adsorbed on a
pentacoordinated AlV Lewis site (Fig. 2).

Figure 3. (a) Macro site on the (100) surface of -alumina


and its schematic representation. Gray: aluminum; red:
oxygen. (b) Simplified network representation of the
kinetic model for isopropanol dehydration on alumina,
including the effect of the inhibition of the reaction by
water affecting only the formation of propene.

4. Conclusions
Comparison between experimental data and multi-
Figure 2. (a) Enthalpic diagram for the formation of
propene from isopropanol on the AlVa site of the - scale modeling shows that dehydration reactions to
Al2O3(100) surface. First transition state structures for olefin and ether may involve the same active site, an
the: (b) E1 pathway, (c) E1cb partway, (d) E2 pathway. AlV Lewis acid site located on the lateral (100)
Grey: Al, Red: O, Blue: C, yellow: H. surface of the alumina grains. As a consequence, the
various paths leading to olefin and ether during
Computed activation enthalpies for the formation of alcohol dehydration cannot be treated as
propene and diisopropylether are close (125 kJ.mol- independent and parallel reactions, but interact
1
/E2 mechanism, and 112 kJ.mol-1/SN2 mechanism, following a complex interplay on a unique site.
respectively); they match very well the experimental Selectivity is thus mainly governed by experimental
values (128 5 and 118 5 kJ.mol-1, respectively). reaction conditions, and not by the action of distinct
Calculations show that the selectivity toward one sites on the catalyst.
product or the other is thus more entropy- than
enthalpy- driven (rS = 8 and 36 J.K-1.mol-1, References
respectively). [1] J. F. DeWilde, H. Chiang, D. A. Hickman, C. R. Ho, A.
Bhan ACS Catalysis 2013, 3, 798.
A microkinetic model was established according to [2] T. K. Phung, A. Lagazzo, M. . Rivero Crespo, V. Snchez
the molecular modeling results, which also included Escribano, G. Busca J. Catal. 2014, 311, 102.
the ether decomposition back to propene and [3] J. H. Kwak, R. Rousseau, D. Mei, C. H. F. Peden, J. Szanyi
isopropanol, occurring at higher temperatures and ChemCatChem 2011, 3, 1557; M. A. Christiansen, G.
also favored on the same active site, as well as water Mpourmpakis, D. G. Vlachos ACS Catalysis 2013, 3, 1965.
[4] K. Larmier, A. Nicolle, C. Chizallet, N. Cadran, S. Maury,
inhibition as proposed by DeWilde et al.[1] (Fig. 3). A.-F. Lamic-Humblot, E. Marceau, H. Lauron-Pernot ACS
The simulated evolution of partial pressures with Catalysis 2016, 6, 1905; K. Larmier, C. Chizallet, N. Cadran, S.
contact time and of selectivity with conversion (Fig. Maury, J. Abboud, A.-F. Lamic-Humblot, E. Marceau, H.
1, solid line) successfully reproduces the Lauron-Pernot ACS Catalysis 2015, 5, 4423.
[5] M. Digne, P. Sautet, P. Raybaud, P. Euzen, H. Toulhoat J.
experimental data on the three materials. Catal. 2004, 226, 54.

Anda mungkin juga menyukai