You are on page 1of 16

Article

pubs.acs.org/IECR

Kinetic Modeling of Propane Oxidative Dehydrogenation over VOx/γ-


Al2O3 Catalysts in the Chemical Reactor Engineering Center Riser
Reactor Simulator
S. Al-Ghamdi,†,‡ J. Moreira,† and H. de Lasa*,†

Chemical Reactor Engineering Centre, University of Western Ontario, London, Ontario, Canada

Research & Development Center, Saudi Aramco Oil Company, Dhahran, Saudi Arabia
*
S Supporting Information

ABSTRACT: This study reports kinetic modeling of propane oxidative dehydrogenation (ODH) employing a new VOx/γ-
Al2O3 catalyst especially designed for propane ODH with a controlled acidity. This catalyst is prepared with different vanadium
loadings (5−10 wt %). Kinetic experiments are carried out under an oxygen-free atmosphere in the Chemical Reactor
Engineering Center fluidized bed riser simulator at 475−550 °C and atmospheric pressure. Successive-injection propane ODH
experiments (without catalyst regeneration) over partially reduced catalysts show good propane conversions (11.73%-15.11%)
and promising propylene selectivity (67.65−85.89%). Regarding propylene selectivity, it increases while that for COx decreases as
the catalyst degree of reduction augments with the consecutive propane injections. This suggests that a controlled degree of
catalyst reduction is needed for high propylene selectivity. Under such oxygen-free conditions, the lattice oxygen of the catalyst is
consumed via the ODH reaction. On the basis of the data obtained, a kinetic model is proposed. In this model, reaction rates are
related to the degree of catalyst reduction using an exponential decay function. The kinetic and decay model parameters are
estimated using nonlinear regression analysis. Activation energies and Arrhenius pre-exponential constants are calculated with
their respective confidence intervals. The proposed parallel-series kinetic model satisfactorily predicts the ODH reaction of
propane under the selected reaction conditions.

1. INTRODUCTION studies have been based on experiments involving cofeeding of


hydrocarbon species and molecular oxygen to the reaction unit.
With the increasing worldwide demand for light olefins, the
However, only a few studies among the above-mentioned
conversion of light paraffins, ethane, propane, and butane, to
research have considered propane ODH in the absence of gas
their corresponding olefins is a topic of current commercial phase oxygen, and only a few of them reported data on kinetic
interest.1−20 Light olefins, such as ethylene and propylene are investigations.45,50,52,58−65 In all these studies, the proposed
important raw materials for the synthesis of polymers and kinetic models involved hydrocarbon molecules reacting with
various petrochemical products. Oxidative dehydrogenation the lattice oxygen of the supported metal oxides, giving olefins
(ODH) of light paraffins has been the most investigated route and H2O. The reduced metal oxides are reoxidized by gas phase
for olefin production, overcoming many limitations of the molecular oxygen to regenerate the active metal oxides and
current olefin production processes, namely, high energy replenish the lattice oxygen.
requirements, coke formation, and thermodynamic con- Moreover, there is limited information that deals with light
straints.21−27 However, the low selectivity toward the desired olefins ODH in dense fluidized beds, risers, or downers
olefin products is the main problem of light paraffin ODH reactors. These studies are being confined mainly to butane
process preventing it from an industrial application.2,28−32 ODH.35,66−68 There is also the motivation of using fluidized-
Light paraffin oxidative dehydrogenation involves a complex bed reactors in light paraffins ODH. This is due to their
reaction network. This reaction network involves competitive controlled isothermal conditions, uniform residence time
reactions taking place simultaneously. In such system, oxygen distributions, and the absence of mass transfer limitations.
supply plays an important role. This is the case as the reaction Hence, one can circumvent with these units, the potential issues
rates for both olefin production and reactants/products with nonisothermal conditions in fixed-bed catalytic reactors.
combustion are strongly influenced by oxygen concentration. Moreover, fluidized reactors with periodic catalyst reoxida-
tion offer the means that enables the transport of the reduced
Thus, optimized oxygen concentration is required in order to
catalyst from the oxidative dehydrogenation reaction unit to the
maximize selectivity toward the desired olefin products.
Kinetic studies for short chain alkane (ethane, propane, and
butane) ODH have been the subject of a significant number of Special Issue: Alirio Rodrigues Festschrift
papers in the open literature.14,16,30,33−57 These contributions Received: December 1, 2013
use different kinetic models such as the Langmuir−Hinshel- Revised: January 29, 2014
wood, the Eley−Rideal, or the Mars van Krevelen models. It is Accepted: February 3, 2014
important to mention that the proposed kinetic models in these

© XXXX American Chemical Society A dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

reoxidation unit. Thus, the operation of twin reactors, one for 3. KINETIC MODELING
the ODH reaction and one for catalyst regeneration appears to Propane ODH involves a network of consecutive and parallel
be a requirement for the implementation of this technology at reactions, namely, the ODH of propane, the undesired
an industrial scale. In this regard, ethane ODH over VOx/γ- combustion of propane feed, and the secondary combustion
Al2O3 has been successfully tested in the Chemical Reactor of propylene product. These reactions can be described
Engineering Center (CREC) riser simulator in the absence of stochiometrically as
molecular gas oxygen. Promising ethane conversions and
ethylene selectivities have already been published.69,70 Thus, 1
C3H8 + O2 → C3H6 + H 2O
it is in our interest and given the significant prospects for 2 (1)
industrial applications to extend our research and explore
1
propane ODH in the CREC riser simulator over VOx/γ-Al2O3 C3H8 + (3x + 4)O2 → 3COx + 4H 2O
catalysts in an oxygen-free atmosphere. 2 (2)
Given the above-described facts, the goal of this paper is to 1
establish propane ODH kinetics over a new VOx/γ-Al2O3 C3H6 + (3x + 3)O2 → 3COx + 3H 2O
2 (3)
fludizable catalyst for propane ODH in the absence of gas
phase oxygen. To accomplish this, a kinetic model in which with the latter two reactions 2 and 3 being the limiting ones for
reaction rates are related to the degree of catalyst reduction propylene selectivity and yield during propane ODH.
using an exponential decay function is proposed. It is our On the basis of the observed dependence between propane
understanding that such propane ODH kinetics in the absence conversion and product selectivities for propane ODH, a
of molecular oxygen gas together with a very selective parallel-series “triangular” reaction network is proposed in this
vanadium-based catalyst and the demonstration of its perform- study as described in Figure 1.
ance in a fluidized bed laboratory reactor has not been reported
before.

2. EXPERIMENTAL SECTION
2.1. Catalyst Preparation and Characterization. The
VOx/γ-Al2O3 catalysts used in this investigation were prepared
by wet saturation-impregnation of a commercial porous γ-Al2O3
support with (SASOL, Catalox SCCa 5/200), with an aqueous
solution of NH4VO3 (Sigma Aldrich, 99%) and oxalic acid
(Sigma Aldrich, 99%) in a 1:2 weight ratio and at a pH of ∼2.
The prepared samples were characterized by BET surface area,
temperature programmed reduction (TPR), X-ray diffraction Figure 1. Proposed reaction network for propane ODH over VOx/γ-
(XRD), and NH3 temperature programmed oxidation (NH3- Al2O3 catalysts in a CREC riser simulator.
TPD), O2 pulse chemisorption, and laser Raman spectroscopy
(LRS). A detailed description of the catalyst preparation and
characterization techniques used are given in a recent The proposed reaction network assumes that propane feed
publication by Al-Ghamdi et al.71 For the sake of avoiding reacts with lattice oxygen promoting two parallel reactions. One
already reported information, the specifics on the preparation of the possible reactions is the formation of the desired
and characterization techniques will not be given in the present propylene product via ODH, with a rate constant k1.
manuscript. Alternatively, propane can be competitively converted forming
2.2. Propane ODH Reaction Studies. The catalytic combustion products (COx), with a rate constant k2. Finally,
propane oxidative dehydrogenation experiments over various the formed propylene may also follow a secondary reaction step
supported vanadium oxide catalysts were established at the leading to the formation of combustion products (COx), with
CREC using the riser simulator. The CREC riser simulator is a rate constant k3.
bench-scale mini-fluidized bed reactor, invented by de Lasa.72 3.1. Kinetic Model Development. Regarding kinetic
This unit is specially designed for catalyst evaluation and kinetic model development and in cases such as the one of the
studies under fluidized bed (riser/downer) reactor conditions. present study, in which propane ODH is carried out in the
Details of the CREC riser simulator as well as the experimental absence of gas phase oxygen, it can be postulated that oxygen
procedure and setup have already been reported elsewhere by from the catalyst lattice is consumed via propane ODH
Al-Ghamdi et al.70 reactions only. Under such conditions, the catalyst oxygen
Thermal and catalytic experiments of propane ODH were content has to be included in the kinetic model. Moreover, an
carried out at four temperatures (475, 500, 525, and 550 °C), allowance must be made for the catalyst’s time on stream or
one catalyst loading (0.76 g) and four contact times (5, 10, 15, catalyst history in order to achieve a kinetic model capable of
and 20 s). All catalytic runs were repeated at three times to describing all transient observations during the course of the
ensure reproducibility of the experimental results. The carbon ODH reaction.
mass balance closures, which considered all carbon-containing To accomplish this, the catalyst oxygen content can be
products such as CO, CO2, CH4, ethylene, ethane, propylene, expressed using a time dependent degree of oxidation (%
propane, and carbon deposited over the catalyst were in the ± 7 oxidation). This time-dependent degree of oxidation can be
% range. For each catalyst, 10 sequential propane injections defined as the ratio of oxygen content left after the ODH
were used at a particular temperature and contact time.. On the reaction over the original oxygen content of the catalyst before
basis of the inlet and outlet concentrations, the conversion, the ODH run. The catalyst degree of oxidation is expected to
selectivity, and yield were calculated. decrease during the ODH reaction cycle. This is especially true,
B dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

as in the present study, if one does not allow catalyst C3H8 − [S2 ](s) + (4 + 3x)[S1O](s)
reoxidation, or replenishment of the lattice oxygen by gas-phase k2
molecular oxygen in between ODH reaction cycles. → 3COx (g) + 4H 2O(g) + (4 + 3x)[S1R ](s)
A possible approach to kinetic modeling based on the
+ [S2 ](s) (8)
available lattice oxygen during oxygen-free ODH reactions is
the use of an exponential decay function based on converted
propane. This type of activity decay function is similar to the (v) Formation of COx from propene (r3) via further reaction
one initially proposed by de Lasa and Al-Khattaf73 for FCC of adsorbed propene with lattice oxygen in an oxidized
catalytic activity decay, Hossain et al.74 for ethylbenzene site [S1O]:
dehydrogenation to styrene, and recently by Al-Ghamdi et al75
for ethane ODH reaction, and it can be written as follows: C3H6 − [S2 ](s) + (3 + 3x)[S1O](s)
→ 3COx (g) + 3H 2O(g) + (3 + 3x)[S1R ](s)
ϕ = exp[−λ(XC3H8)] (4)
+ [S2 ](s) (9)
where ϕ is the catalyst’s degree of oxidation, λ is a constant, and
X is the propane conversion. The main advantage of this (vi) Reoxidation of the reduced metallic center by molecular
function is that it accounts for the effects of reaction conditions oxygen:
(temperature, concentration, and contact time) on the catalyst
extent of oxidation. O2 (g) + 2[S1R ](s) → [S1O](s) + [S1O](s) (10)
Furthermore, the choice of the kinetic model for propane On the basis of a MVK mechanism, the rate equations
ODH is of paramount importance. With this in mind, a Mars corresponding to each surface reaction in the above steps, (i.e.,
van Krevelen (MVK) mechanism is considered in the present r1, r2, and r3) are expressed as follows:
study. This model is generally accepted for oxidative
dehydrogenation reactions over vanadia catalysts involving r1 = k1θC3H8(1 − β) (11)
VOx lattice oxygen species. On the basis of a MVK mechanism
established in the context of two catalyst sites for propane r2 = k 2θC3H8(1 − β) (12)
ODH, the catalyst is reduced by the reaction of the adsorbed
propane with catalyst lattice oxygen to produce propylene r3 = k 3θC3H6(1 − β) (13)
molecules and carbon oxides. Moreover, the adsorbed
propylene product on the catalyst surface can further react where ri is the reaction rate (mol/g·sec), ki is the reaction rate
with lattice oxygen producing carbon oxides. Gas phase oxygen constant (mol/g·sec), θi is the surface coverage of adsorbed
can be intermittently introduced following several propane species “i”, with β + γ = 1, in which β represents the reduced
injections to replenish the lattice oxygen by reoxidation of the vanadium sites and γ = 1 − β represents the oxidized vanadium
catalyst. sites (site 1 type), and θC3H8 + θC3H6 + θv = 1 describes the
As a result, the following six-step mechanism is proposed as alumina based sites (site 2 type).
shown below, with [S1O] representing lattice oxygen in an Thus, and as a result, the surface coverage θi is given by
oxidized site-1, [S1R] denoting a surface oxygen vacancy in a
K C3H8CC3H8
reduced site-1, and [S2] representing alumina-based sites θC3H8 =
(i) Propane adsorption on a site type-2 [S2] on the catalyst (1 + K C3H8CC3H8 + K C3H6CC3H6) (14)
surface:
K C3H6CC3H6
K C3H8 θC3H6 =
C3H8(g) + [S2 ](s) ←⎯⎯→ C3H8 − [S2 ](s) (5) (1 + K C3H8CC3H8 + K C3H6CC3H6) (15)

where Ki is the adsorption constant (cm3/mol) of species “i”


(ii) Formation of propylene (r1) via the surface reaction of
and Ci is the concentration (mol/cm3) of species “i”. Moreover,
adsorbed propane with lattice oxygen in an oxidized site given the proposed decay model for the catalyst’s degree of
[S1O] via H-abstraction: oxidation (ϕ) in eq 4, β can be related to ϕ as
C3H8 − [S2 ](s) + 2[S1O](s) (1 − β) = φ (16)
k1
→ C3H6 − [S2 ](s) + H 2O(g) + [S1O](s) A substitution of eqs 14 to 16 into eqs. 11 to 13 gives the
following rate equations based on the proposed reaction
+ [S1R ](s) (6) network in Figure 1:
k1K C3H8CC3H8
(iii) Desorption of adsorbed propylene from a reduced site: r1 = exp[−λ(XC3H8)]
K C3H6
(1 + K C3H8CC3H8 + K C3H6CC3H6)
C3H6 − [S2 ](s) ←⎯⎯→ C3H8(g) + [S2](s) (7) (17)

k 2K C3H8CC3H8
(iv) Formation of COx from propane (r2) via the reaction of r2 = exp[−λ(XC3H8)]
adsorbed propane with lattice oxygen in an oxidized site (1 + K C3H8CC3H8 + K C3H6CC3H6)
[S1O]: (18)

C dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

k 3K C3H6CC3H6 dyi Wc ⎛ MWi VR ⎞


r3 = exp[−λ(XC3H8)] = ⎜ ⎟ri
(1 + K C3H8CC3H8 + K C3H6CC3H6) dt VR ⎝ Whc ⎠ (25)
(19)
This equation is appropriate under the conditions of the
One can notice in eqs 17,18, and 19 that the rate of reaction is present study given that (a) the CREC riser simulator is a
related to the adsorption constants, the catalyst extent of bench-scale isothermal well-mixed batch reactor unit with a
oxidation, and the intrinsic kinetic parameters. high fluid gas recirculation and mixing times in the 20−30 ms
Furthermore, using these rate equations, one can establish scale77 and (b) the catalyst particles operate freely from
the overall rate of consumption and/or formation of chemical intraparticle diffusional transport controls. The second
species. This can be accomplished on the basis of the algebraic condition is met given the small size of catalyst particles (90
addition of reaction rates consistent with the reaction network μm) and the porous structure of the fludizable alumina used.
given in Figure 1. On the basis of this, the following rate By establishing one equation such as eq 25 for each
equations could be established for each chemical species component, one can obtain a set of differential equations to
involved in the propane ODH reaction network: represent the propane ODH network. This set of ordinary
differential equations is obtained by substituting the rate
Rate of Propane Consumption: equation of each component given in eqs 20 to 22 into eq 25
and considering that
rC3H8 = −(r1 + r2) yC H Whc
3 8
−(k1 + k 2)K C3H8CC3H8 CC3H8 =
= exp[−λ(XC3H8)] MWC3H8VR (26)
(1 + K C3H8CC3H8 + K C3H6CC3H6)
and
(20)
yC H Whc
Rate of propylene formation: CC3H6 = 3 6

MWC3H6VR (27)
rC3H6 = r1 − r3
One can obtain
(k1)K C3H8CC3H8 − (k 3)K C3H6CC3H6
= exp[−λ(XC3H8)]
(1 + K C3H8CC3H8 + K C3H6CC3H6) for propane:
(21)
dyC H −Wc ⎛ MWC3H8 VR ⎞
Rate of COx formation: 3 8
= ⎜ ⎟
dt VR ⎝ Whc ⎠
rCOx = r2 + r3 (k1 + k 2)αK C3H8yC H
3 8
(k 2)K C3H8CC3H8 + (k 3)K C3H6CC3H6 ×
= exp[−λ(XC3H8)] (1 + αK C3H8yC H + βK C3H6yC H )
3 8 3 6
(1 + K C3H8CC3H8 + K C3H6CC3H6)
(22)
× exp[−λ(XC3H8)] (28)
3.2. Kinetic Modeling in the CREC Riser Simulator. for propylene:
Kinetic modeling in the CREC riser simulator requires species
balances. Such species balance equations can be established on
dyC H Wc ⎛ MWC3H6VR ⎞
3 6
= ⎜ ⎟
the basis of a well mixed isothermal batch-scale reactor.76,77 dt VR ⎝ Whc ⎠
Thus, the following CREC riser simulator equation can be
considered: k1αK C3H8yC H − k 3βK C3H6yC H
3 8 3 6
×
VR dCi (1 + αK C3H8yC H + βK C3H6yC H )
3 8 3 6
ri =
Wc dt (23) × exp[−λ(XC3H8)] (29)
where VR is the volume of riser simulator (cm3), Wc is the
for COx:
weight of the catalyst (g), Ci is the concentration of species “i”
(mol/cm3), and t is the time (seconds). dyCO Wc ⎛ MWCOxVR ⎞
The concentration of any species (Ci) is related to its weight
x
= ⎜ ⎟
fraction (yi) as dt VR ⎝ Whc ⎠
k 2αK C3H8yC H + k 3βK C3H6yC H
yW
i hc 3 8 3 6
Ci = ×
MWi VR (24) (1 + αK C3H8yC H + βK C3H6yC H )
3 8 3 6

where Whc is the total mass of hydrocarbons injected into the × exp[−λ(XC3H8)] (30)
riser simulator (g), MWi is the molecular weight of ith
component (g/mol), and VR is the riser simulator volume where α = (Whc)/(MWC3H8VR) and β = (Whc)/(MWC3H6VR)
(cm3). The system of ODE given in eqs 28 to 30 can be solved with
After replacing eq 24 into eq 23 and after the required a carefully selected set of initial conditions. This allows
algebraic steps, the CREC riser simulator design equation is describing the mass fraction changes of various chemical
obtained in terms of weight fractions chemical species: species (e.g., propane, propylene, and COx) with reaction time.
D dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 1. Characteristics of Alumina Support and Supported Vanadia Catalyst


V contenta SBET (m2/ H2 uptake (cm3 % surface density NH3 uptake (cm3 Lewis/Brønsted % dispersionc
sample (wt %) g) STP/g) reductionb (V/nm2) STP/g) ratio (O/V)
γ-Al2O3 0 203.52 0.54 16.26
(activated)
5% V/γ-Al2O3 5 188.64 0.45 81.28 3.1 10.94 1.8 63.26
7% V/γ-Al2O3 7 179.27 0.46 83.06 4.6 11.50 4.2 59.81
10% V/γ-Al2O3 10 153.37 0.39 99.18 7.7 13.02 5.1 43.25
a
V content, in wt % was determined by ICP. bBased on V2O5 + 2H2 → V2O3 + 2H2O. cDispersion = fraction of vanadium atoms at the surface,
assuming Oads/Vsurf = 1

Furthermore, to obtain the intrinsic kinetic parameters NH3-TPD and pyridine FTIR analysis showed that both the
(activation energies (Ei) and pre-exponential factors (kio)), number and nature of the acid sites change with vanadium
the temperature dependence of each (ki) parameter in eqs 28 to loading. The surface vanadia species (VOx) cover the exposed
30 can be represented using the Arrhenius’ equation as follows: alumina surface Lewis acid sites, resulting in lower acidity than
that of the bare support. Moreover, the number of surface
⎡ −E ⎛ 1 1 ⎞⎤
ki = kio exp⎢ i ⎜ − ⎟⎥ Brønsted acid sites increased with increasing surface VOx. This
⎢⎣ R ⎝ T Tm ⎠⎥⎦ (31) indicated that polymeric surface VOx species possess more
Brønsted acid character than the isolated surface VOx species.
where koi is the pre-exponential factor (mol/g·sec) expressing X-ray diffraction (XRD) demonstrated the absence of V2O5
ODH vanadium oxidize catalyst reactivity at reaction time zero bulk surface species and the high dispersion of VOx on the
and T = Tm, Ei is the activation energy (kJ/mol), R is the support surface. No evidence of AlV3O9 was detected in any of
universal gas constant, and Tm is the average temperature (K). the samples suggesting that the reaction between VOx and
To reduce the cross-correlation between the pre-exponential Al2O3 is negligible during the treatment at 600 °C.
factors koi and the activation energies Ei, the ki constants were Moreover, XRD analysis on spent catalysts after 10
reparameterized, as shown in eq 31. This was accomplished by consecutive ODH propane injections and at the most severe
centering the reaction temperature at an average value of Tm = reaction conditions showed no diffraction patterns. This result
512.5 °C. This temperature corresponds to the average reaction was assigned to the structural changes other than those
temperatures used in the present study as recommended by occurring in the bare γ-Al2O3 support. This can also be viewed
Hossain and de Lasa.78 as an indication of the high catalyst stability under the reaction
In the same way and on the basis of adsorption conditions.
thermodynamics, one can relate the adsorption constant Ki 4.2. Propane ODH in CREC Riser Simulator. Table 2
with the reaction temperature T, as reports propane conversion and product selectivities for the
⎡ −ΔH ⎛ 1 successive-injection propane ODH experiments over the
1 ⎞⎤
K i = K io exp⎢ i
⎜ − ⎟⎥ various VOx/γ-Al2O3 catalysts at various contact times and
⎢⎣ R ⎝ T Tm ⎠⎥⎦ (32) different temperatures. One should also notice that the only
identifiable carbon-containing products other than propylene
where −ΔHi is the heat of adsorption (kJ/mol) and is the koi were CO, CO2, CH4, C2H6, and C2H4.
pre-exponential factor (cm3/mol). The substitution of eq 32 On the basis of the product distribution results shown in
and eq 33 into eqs 29 to 31 gives a new set of ordinary Table 2, the variation of product selectivities with propane
differential equations with the intrinsic kinetic parameters (koi , conversion is shown in Figure 2, for the various VOx/γ-Al2O3
Ei, Koi and −ΔHi) to be estimated. catalysts. It is apparent from Figure 2 that the propene
selectivity is inversely related to the conversion for each
4. RESULTS AND DISCUSSION catalyst.
4.1. Catalyst Characterization. Catalyst characterization The decrease in propylene selectivity with a corresponding
results are summarized in Table 1. Additional information increase in COx selectivities shows that propylene is the
regarding catalyst characterization can be found in our recent primary reaction product of the propane ODH while COx is the
publication by Al-Ghamdi et al.71 BET surface area analysis of secondary product of the consecutive oxidation of propane feed
the prepared VOx/γ-Al2O3 catalysts showed a moderate and propylene product. On the basis of these results, a parallel-
decrease in the total surface area of the catalyst upon vanadium series “triangular” reaction network with “direct alkane
loading on the calcined γ-Al2O3 support. This decrease of the combustion” together with “alkene formation” and later “alkene
surface area was due to the plugging of some of the support combustion” could be proposed for propane ODH as given in
pores by vanadium species. Figure 1. Further discussion about the reaction network is given
H2-TPR showed that all VOx/γ-Al2O3 catalysts exhibited a in the Kinetic Modeling section.
single low temperature reduction peak at around 450 °C which Figure 3 reports the various gas species conversions/
was due to the reduction of isolated and two-dimensional selectivities as a function of a propane injection for two sets
structures of surface vanadia. Moreover, the reducibility (% of propane ODH experiments over the various VOx/γ-Al2O3
reduction) of the surface VOx species increased with surface catalysts. Figure 3a,b describes successive-propane injections
VOx surface density in the submonolayer region. during ODH runs. The catalyst is not regenerated in between
O2-chemisorption and Raman spectroscopy demonstrated propane injections (pulses). It can be observed in Figure 3a
that the structure and dispersion of the VOx surface species that the conversion of propane is highest after the first injection
depend on their surface density. The surface vanadia species for all catalysts, and then decreases with the increasing number
becomes more polymerized with increasing vanadium loading. of propane injections.
E dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 2. Conversion and Product Distribution Results for Sequential-Injections Propane ODH Experiments over Various VOx/
γ-Al2O3 Catalystsa
selectivity (%)
o
catalyst time (s) T ( C) XC3H8 (%) C3H6 COx CH4 C2H6 C2H4 C3H6 yield (%)
5% VOx/γ-Al2O3 5 475 2.35 70.89 86.49 2.76 0.24 0.70 1.63
500 3.08 73.37 86.26 4.73 0.60 1.30 2.18
525 3.83 75.44 88.01 6.14 1.51 2.26 2.85
550 4.08 76.38 89.66 6.30 1.54 3.78 3.13
7% VOx/γ-Al2O3 475 3.24 60.73 35.81 2.57 0.25 0.64 1.62
500 3.74 64.75 29.00 4.45 0.65 1.16 2.14
525 4.60 65.29 24.36 6.31 1.76 2.28 2.63
550 5.06 66.20 20.01 7.12 2.43 4.24 3.18
10% VOx/γ-Al2O3 475 3.73 55.12 41.52 2.48 0.27 0.61 1.48
500 4.22 57.24 36.74 4.27 0.71 1.04 2.01
525 5.26 58.79 30.29 6.60 2.06 2.26 2.47
550 6.16 60.64 23.79 7.83 3.18 4.56 3.49
5% VOx/γ-Al2O3 10 475 3.29 76.86 89.88 2.97 0.27 0.67 2.50
500 4.59 77.53 89.96 4.20 0.69 1.06 3.46
525 6.32 79.92 91.09 5.67 2.10 2.03 5.03
550 6.08 81.24 91.86 5.49 2.20 2.76 4.98
7% VOx/γ-Al2O3 475 4.18 67.20 29.30 2.61 0.30 0.59 2.59
500 5.60 68.94 24.78 4.29 0.89 1.11 3.53
525 6.55 70.60 19.46 5.79 2.20 1.94 4.34
550 8.41 71.49 15.64 6.60 2.98 3.30 5.90
10% VOx/γ-Al2O3 475 4.67 61.99 34.59 2.49 0.33 0.59 2.57
500 6.27 63.43 30.05 4.44 1.03 1.06 3.42
525 7.24 64.75 25.05 6.04 2.28 1.88 4.10
550 10.50 66.18 18.79 7.51 3.68 3.84 6.64
5% VOx/γ-Al2O3 15 475 4.19 80.35 91.53 2.97 0.29 0.67 3.35
500 5.77 81.89 91.89 4.98 0.94 1.12 4.71
525 7.99 83.28 92.90 5.58 2.60 1.99 6.67
550 9.12 84.52 93.88 5.05 3.30 2.44 7.73
7% VOx/γ-Al2O3 475 5.14 67.68 28.76 2.64 0.32 0.61 3.29
500 6.89 71.58 22.00 4.37 1.01 1.03 4.69
525 9.77 71.98 19.11 5.00 2.31 1.61 6.78
550 11.21 74.46 11.67 6.73 3.92 3.22 8.36
10% VOx/γ-Al2O3 475 5.90 61.67 34.82 2.60 0.34 0.57 3.32
500 7.68 64.64 29.27 4.13 1.03 0.92 4.52
525 10.79 65.78 25.89 4.85 2.07 1.41 6.56
550 12.93 66.81 16.13 8.53 4.75 3.78 8.53
5% VOx/γ-Al2O3 20 475 5.38 84.42 94.45 2.94 0.31 0.61 4.53
500 7.32 84.70 95.22 4.15 0.89 0.96 6.21
525 8.88 85.79 95.52 4.84 2.30 1.75 7.64
550 11.73 85.94 96.90 5.41 3.62 1.55 10.11
7% VOx/γ-Al2O3 475 6.42 71.20 25.19 2.73 0.34 0.55 4.36
500 8.11 73.09 21.03 4.02 0.99 0.87 5.82
525 11.88 73.79 17.51 4.78 2.40 1.51 8.45
550 13.36 75.34 10.43 7.21 4.44 2.58 10.06
10% VOx/γ-Al2O3 475 7.16 64.13 32.45 2.53 0.36 0.53 4.24
500 9.14 65.73 28.66 3.72 1.07 0.82 5.70
525 13.41 67.29 23.58 5.21 2.48 1.43 8.32
550 15.05 67.77 15.01 8.72 4.99 3.51 10.16
a
Reported values are the average of 10 consecutive C3H8 injections.

As no molecular oxygen is fed with the injection, the catalyst that the decrease of propane conversion over VOx/γ-Al2O3
is believed to supply all the oxygen for the ODH reaction. catalysts mainly resulted from the consumption of the reactive
Following the first injection, propane conversion declined lattice oxygen species present on the catalyst with propane
drastically from 58.9, 76.5, and 81.9 to 12.5, 14.4 and 16.4%. injections. Moreover, it can be seen for all catalysts that the first
Following this, there was a further gradual decrease to 11.5, injection also has very low propylene selectivity and high COx
12.4, and 13.6% with subsequent propane injections over the selectivity.
5%VOx/γ-Al2O3, 7%VOx/γ-Al2O3, and 10%VOx/γ-Al2O3 cata- Fully oxidized VOx/γ-Al2O3 catalysts are considered non-
lysts samples, respectively. Therefore, it is plausible to suggest selective for the ODH reaction as seen from the results
F dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 2. Propylene selectivities (filled symbols) and COx selectivities (empty symbols) as a function of C3H8 conversion (C3H8 injected = 10 mL,
contact time = 20 s, catalyst loaded = 0.76 g). Error bars correspond to standard deviation of three trials (data reported for 10 consecutive pulses
without catalyst regeneration between the injections).

obtained at the first injection. When the catalyst is in a fully These findings demonstrate that in consecutive ODH
oxidized state, there is a higher population of nonselective reactions without catalyst regeneration in between (oxygen-
surface oxygen species which is more likely to promote the free environment), propane can be converted selectively to
deep oxidation of hydrocarbons to carbon oxides. These propylene. This occurs due to the lattice oxygen involvement in
nonselective oxygen species could be formed due to either the the ODH reaction mechanism. One can also notice that
loosely bound lattice oxygen or the adsorbed oxygen from the following the first ethane injection (first pulse), the catalyst
regeneration step which could have remained on the VOx/γ- tested displays a stable performance in terms of ethane
Al2O3 catalyst surface. conversion and ethylene selectivity even after 10 successive
However, it can be noticed in Figure 3b that following the reduction runs.
first propane injection, the selectivities for propylene increase On the other hand, Figure 3c,d reports a second set of
experiments involving repeated ODH reaction-regeneration
with propane injections showing stable values at around 92%,
cycles at 550 °C over the various VOx/ γ-Al2O3 catalysts.
82.2%, and 76.1% over the 5%VOx/γ-Al2O3, 7%VOx/γ-Al2O3
Catalyst regeneration was conducted by flowing air at 550 °C
and 10%VOx/γ-Al2O3 catalysts samples, respectively. On the over the catalyst after each propane injection (pulse). Results
other hand, selectivities for COx decrease with propane show that propane conversion and product selectivities
injections and this seems to indicate that a certain degree of (equivalent to a per-pass conversion and selectivity in a
catalyst reduction is required in order to obtain good propylene continuous riser/downer unit) remain stable during the
selectivity. In this regard, a selective catalyst surface would be different runs. It can be observed in Figure 3c,d that propane
obtained only after the adsorbed oxygen had been consumed by conversion remains at the 60, 73.2, and 80.1% values over the
the first propane injection. This is evident from the subsequent 5%VOx/γ-Al2O3, 7%VOx/γ-Al2O3, and 10%VOx/γ-Al2O3 cata-
propane injections where the selectivity to propylene increases lysts samples, respectively. On the other hand, very low
with the number of propane injections while that for propylene selectivities in the 2.1, 0.93, and 0.57% levels are
combustion products (COx) decreases. observed over the 5%VOx/γ-Al2O3, 7%VOx/γ-Al2O3, and 10%
G dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 3. Propane conversion, propylene, and COx selectivities over various VOx/γ-Al2O3 catalysts in the CREC riser simulator. (T = 550 °C, C3H8
injected = 10 mL, reaction time = 20 s, catalyst loaded = 0.76g): (a,b) consecutive ODH runs without catalyst reoxidation in between propane
pulses; (c,d) consecutive ODH runs with catalyst reoxidation in between propane pulses. Error bars correspond to standard deviation of three trials.

Table 3. Intrinsic Kinetic Parameters for the Proposed Kinetic Model. Kinetic Parameters Are Reported with Their 95%
Confidence Interval over Various VOx/γ-Al2O3 catalysts
5% VOx/γ-Al2O3 7% VOx/γ-Al2O3 10% VOx/γ-Al2O3
parameter value 95% CI value 95% CI value 95% CI
ko1a 3.17 x10−3 ±3.89 × 10−5 4.82 x10−3 ±1.00 x10−4 6.34 x10−3 ±2.68 x10−4
ko2 9.28 x10−4 ±1.52 x10−5 3.04 x10−3 ±6.76 x10−5 5.91 x10−3 ±2.30 x10−4
ko3 8.52 x10−5 ±6.78 x10−4 2.97 x10−4 ±1.67 x10−3 6.33 x10−4 ±3.06 x10−3
KoC3H8b 76.72 ±0.93 51.14 ±1.01 38.34 ±1.45
KoC3H6 29.95 ±3.65 28.58 ±5.37 26.12 ±7.82
E1c 124.92 ±1.33 115.08 ±0.99 109.42 ±2.52
E2 52.81 ±1.71 51.07 ±1.42 45.58 ±2.34
E3 52.54 ±8.99 52.73 ±10.94 53.75 ±17.55
−ΔHC3H8c 68.18 ±1.41 50.47 ±1.14 40.32 ±2.18
−ΔHC3H6 54.59 ±8.22 56.86 ±10.80 41.60 ±12.34
σd 2.15 × 10−4 2.65 × 10−4 3.90 × 10−4
λe 0.01−0.053 0.017−0.056 0.015−0.047
m 960 960 960
DOF 950 950 950
a
mol gcat−1 s−1. bcm3/mol. ckJ/mol. dσ = (∑(Xexperimental − Xestimated)2/(m − p))1/2, where m is the number of data points, p is the number of model
parameters, and DOF is the degree of freedom. eDeactivation constant (λ) calculated by independent fitting of experimental data.

VOx/γ-Al2O3 catalysts samples, respectively. In addition, COx γ-Al2O3, 7%VOx/ γ-Al2O3, and 10%VOx/γ-Al2O3 catalysts
selectivities remain stable at 88, 90, and 93% over the 5%VOx/ samples, respectively.
H dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 4. Energy diagram for the ODH reaction: (a) propane ODH reaction forming propylene, (b) propane oxidation forming COx, (c) propylene
oxidation forming COx.

Table 4. Standard heat of Reaction for Propane and directly; (b) only a small fraction of the catalyst stream is
Propylene Combustion directed toward the regenerator. This operational strategy can
allow the functioning of industrial scale ODH processes with
reaction ΔHorxn (kJ/mol)
high propylene selectivities, under conditions close to the ones
1 reported in Figure 3a,b.
C3H8 + O2 → C3H6 + H 2O −117.6
2 4.3. Kinetic Parameters Estimation. The postulated rate
7 expressions in eqs 28 to eq 30 are nonlinear with respect to
C3H8 + O2 → 3CO + 4H 2O −1195.1
2 their parameters. This is due to the fact that the parameters
C3H8 + 5O2 → 3CO2 + 4H 2O −2043.9 appear both in the numerator and in the denominator of the
C3H6 + 3O2 → 3CO + 3H 2O −1077.5 rate expressions such as in the case of the adsorption constant
9
(Ki). Therefore, the estimation of the 10 parameters (ko1, ko2, ko3,
C3H6 +
2
O2 → 3CO2 + 3H 2O −1926.4 E1, E2, E3, KoC3H8, KoC3H6, −ΔHC3H8 and −ΔHC3H6) was developed
using nonlinear least-squares regression. The MATLAB routine
“LSQCURVEFIT” was used for the regression analysis. The
Regarding propylene and COx selectivities during repeated numerical integration of the system of ODE given in eqs 29 to
ODH reaction−regeneration cycles, they are in clear contrast 31 and the determination of the 95% confidence intervals for
with the ones reported in Figure 3a,b for consecutive propane each estimated parameter were performed using the MATLAB
ODH runs. They show that the formation of the CO and CO2 functions “ode113”; and “nlparci”, respectively.
is promoted by gas phase oxygen introduced for the catalysts Experimental data at different reaction temperatures were
regeneration. These low propylene selectivities and higher COx used to evaluate reaction rate parameters. Moreover, an
selectivities obtained at any given propane conversion during Arrhenius type of temperature-dependence function was used
the repeated cycles of single injection ODH experiments can be to express the specific rate constants. This was done by taking
related to both the amount and type of surface oxygen species into account the temperature dependence of the experimental
available on the fresh catalyst surface. Excess amounts of surface data as shown in eq 32.
oxygen as well as the existence of nonselective oxygen species Moreover, the deactivation constant “λ” given in eq 4 was
on the fresh catalyst surfaces are believed to favor the obtained through an independent fitting of the propane
conversion of propane and the produced propylene into conversion data for every run (e.g., using the propane
combustion products. conversions obtained for injection 4 at 5 s, 10 s, 15 s, and 20
In view of the above, it can be concluded that fully oxidized s). The ‘λ” parameter was calculated via nonlinear regression.
(fresh) VOx/γ-Al2O3 catalysts are active but not selective for Regression coefficients were, in all cases, in excess of 0.98. The
propane ODH reactions. This fact can be attributed to the obtained values of the deactivation constant “λ” were used as
excess (nonstoichiometric) amount of oxygen species on the inputs in the proposed kinetic model along with all the mass
catalysts surface. These species could be either loosely bound fractions of the reactants/products. This approach was valuable
lattice oxygen or weakly adsorbed oxygen species produced to eliminate the high cross-correlation observed between λ and
from gas-phase O2 during pretreatment and the catalyst the other kinetic parameters
regeneration. Both types of oxygen species may be involved The optimization criteria considered was established having
in the total oxidation of propane and propylene product. the following bounds for the model parameters: (a) rate
Thus, one can conclude that the performance of the various constants and activation energies for each reaction must be
VOx/γ-Al2O3 catalysts of the present study is affected positive, (b) heats of adsorption (ΔHi) must be negative. On
considerably by both ODH operation conditions and catalyst this basis, a minimum sum of squares of errors was calculated:
regeneration. It appears that consecutive reaction injections
(pulses) are a preferred mode of operation for ODH. As a N
result, ODH catalyst regeneration shall only be allowed once SSR = ∑ (xi ,exp − xi ,pred)2
this series of consecutive reaction cycles are completed. i=1 (33)
One can, for instance, envision that this is equivalent to
operating a twin circulating fluidized reactor process (reactor− where xi,exp and xi, pred are the mass fraction of component i
regenerator), where the following occurs: (a) most of the obtained experimentally and predicted by the kinetic model,
catalyst stream leaving the reactor is returned to the ODH unit respectively.
I dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 5. Comparison between experimental data and model predictions () over a 5% VOx/γ-Al2O3 catalyst sample: (a) T = 475 °C; (b) T = 500
°C; (c) T = 525 °C; (d) T = 550 °C. (Data points are reported for the average of 2nd to 10th injections.)

Furthermore, discrimination between possible models was increase with VOx surface density. This suggests that
based on (i) correlation coefficients (R2), (ii) lower SSR (sum polyvanadate surface species, which are more dominant at
of the squares of the residuals) criteria, (iii) lower cross- higher VOx surface densities, are more active than mono-
correlation coefficient (γi), and (iv) smaller individual vanadate species.
confidence intervals for the model parameters. Similarly, the increase in the ko2 and ko3 pre-exponential factors
The values of the 10 estimated parameters along with their with vanadium loading is also in agreement with the
corresponding 95% confidence intervals (CI) are reported in experimental results. It was also found that COx formation
Table 3 for the different VOx/γ-Al2O3 catalysts. It can be rates increase with VOx surface density. This can be attributed
noticed that all the estimated parameters display a reduced and to the excess surface oxygen species available at higher loadings.
acceptable 95% C.I. Moreover, the ability of establishing the 10 This is also consistent with the decreasing propylene selectivity
model parameters is consistent with the high DOF (degree of with VOx surface density as more COx are formed.
freedom) in this analysis with 960 experimental data points for Regarding the relative magnitude of the activation energies
each catalyst sample considering three trials for each run.
obtained as reported in Table 3, it can be noticed that COx
One can also observe that all 10 parameters in Table 3 were
formation activation energies (E2 and E3) are consistently
calculated with reduced spans. In addition, several kinetic
parameter trends regarding the effect of vanadium loading on smaller than E1. On the basis of this observation, one could
the catalytic activity can be observed in Table 3. One can notice argue that the obtained values are consistent with the energy
that the pre-exponential factors (ko1, ko2, and ko3) augment as the diagrams given in Figure 4. One can observe in these diagrams,
vanadium loading is increased. the major difference of the heats of reaction: the COx heat of
In view of these findings, one can consider correlating reaction being much larger than the C3H6 heat of formation via
changes in the ko1, ko2, and ko3 values with the different vanadia ODH. Thus, on this basis, one can also envision the propane
structures present at specific vanadium loadings, though other ODH reaction forming propylene with activation energy larger
factors may also be involved. In this respect, one can also notice than the ones for the reactions leading to COx as reported in
that the increase in ko1 is in agreement with the experimental Table 4. These facts may contribute to explaining the limited
results. Specific initial rates of propane ODH were found to ODH propylene selectivity at higher temperatures.
J dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 6. Comparison between experimental data and model predictions () over a 7% VOx/γ-Al2O3 catalyst sample: (a) T = 475 °C; (b) T = 500
°C; (c) T = 525 °C; (d) T = 550 °C. (Data points are reported for the average of 2nd to 10th injections.)

In addition, further insights into the validity of the proposed related to physicochemical properties of the catalysts under
kinetic model and the estimated kinetic parameters could be study.
obtained. This can be done by comparing the model On the basis discussed above, it can be concluded that the set
predictions for products and reactant mass fractions with of adsorption and kinetic parameters established as well as the
experimental data. This was, in fact, done for all the catalysts of kinetic model developed are adequate to predict propane ODH
this study as reported in Figures 5−7. It can be noted that, reaction rates in the CREC riser simulator in the range of the
within the limits of experimental error, the model predictions operating conditions studied.
compare very well with the experimental data. Thus, this It is important to review the values of the activation energies
comparison validates the adequacy of the proposed reaction obtained in the present study and to compare them with values
model. already reported in the technical literature. Table 6 reports the
Moreover, it can also be inferred from the parity plots activation energies for propane ODH on various supported
reported in Figures 5−7 for model predictions, as compared to VOx catalysts. Although the reported values for the three
the experimental data, that the data is not clustered in propane ODH catalysts of this study are consistently within the
horizontal or vertical lines. Horizontal bands may be the result same range, there is some discrepancy between these energies
of changes in the observed conversion caused by an of activation (± 20 kJ/mol) when compared with other values
independent variable which is not included in the kinetic published in the technical literature. This discrepancy can be
model. On the other hand, vertical lines are an indication of the attributed to the dissimilarity in experimental systems,
kinetic model overparameterization.79,80 operating conditions, proposed ODH kinetic model, and
The adequacy of the estimated parameters was checked by catalysts studied.
analyzing their dependence on each other through the cross- It should be stressed, however, that the values of activation
correlation matrix as shown in Table 5. It can be observed, that energies for the propane ODH of this study are close to the
in most cases, cross-correlation coefficients are below 0.90 with ones reported by Rao and Deo, 200446 using VOx/γ-Al2O3
only three of them surpassing the 0.9 value. Hence, it can be catalysts with similar vanadium loadings. For instance, for the
concluded that the kinetic model as proposed is not propane ODH step, Rao and Deo, 2004 reported activation
overparameterized and that the defined parameters can be energies for propylene formation of 113.6, 100, and 98 kJ/mol
K dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Figure 7. Comparison between experimental data and model predictions () over a 10% VOx/γ-Al2O3 catalyst sample: (a) T = 475 °C; (b) T =
500 °C; (c) T = 525 °C; (d) T = 550 °C. (Data points are reported for the average of 2nd to 10th injections.)

for catalysts with vanadium loadings of 7.5, 10, and 12.5 wt %, support an ODH process using twin circulating fluidized beds
respectively. The values of activation energies observed in this with (a) an ODH reactor with oxygen being supplied from the
study for the same step were 124.9, 115.1, and 82 KJ/mol for 5, catalyst lattice only and (b) a reoxidation reactor where the
7, and 10 wt % catalysts, respectively. catalyst lattice oxygen is replenished. It is anticipated that in this
However, in the case of COx formation, the reported values type of ODH process, only a 1/23 fraction of the catalyst in the
by Rao and Deo, 200446 differ by ±20 kJ/mol from those reactor outlet stream is going to be treated per pass in the
obtained in the context of the present study. This discrepancy reoxidation reactor. This is estimated considering up to 23
between activation energies of COx formation can be attributed injections are required in the CREC riser simulator until
to both the different reaction considered by Rao and Deo, oxygen is fully depleted from the VOx/γ-Al2O3 catalysts.71 The
2004, where oxygen is being cofed with propane, and the rest or most of the catalyst in the ODH reactor outlet stream is
different reaction network paths involved. Rao and Deo, 2004, going to be partially cooled and recycled back directly to the
assumed that CO and CO2 are formed from propylene instead reactor input stream. It is expected that such a twin and
of following the the parallel-series “triangular” reaction network integrated ODH fluidized bed process will yield higher olefin
of the present study. This proposed “triangular” reaction selectivities as reported in the present study.
network of the present study is reinforced by the fact that COx
species are observed even at very low propane conversions. 5. CONCLUSIONS
Thus, a direct COx formation step from propane is required. The following key points can be considered as the main
As a result, on the basis of the experimental evidence conclusions of this study:
gathered and on the adequacy of the parallel-series “triangular” 1. Propane was converted to propylene via ODH reaction
kinetic model proposed, the following can be considered as over VOx/γ-Al2O3 catalysts utilizing the lattice oxygen of
reported in Figure 1: (i) There is an alkene formation step, (ii) the catalysts. Good propane conversion (11.73%−
there is a competitive direct alkane combustion step, and (iii) 15.11%) and promising propylene selectivity (67.65−
once alkene is formed, there is further alkene combustion. 85.89%) were achieved in successive-injection propane
In summary, the catalytic runs of the present study together ODH experiments (without catalyst regeneration) in the
with the phenomenologically based kinetic model developed CREC riser simulator at 475−550 °C.
L dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX
Industrial & Engineering Chemistry Research Article

Table 5. Cross-Correlation Coefficients for the Kinetic Model Optimized Parameters over Various VOx/ γ-Al2O3 Catalyst
Samples
ko1 ko2 ko3 KoC3H8 KoC3H6 E1 E2 E3 −ΔHC3H8 −ΔHC3H6
(a) 5% VOx/γ-Al2O3 Catalyst Sample
ko1 1.000
ko2 0.657 1.000
ko3 0.150 −0.617 1.000
KoC3H8 −0.975 −0.803 0.061 1.000
KoC3H6 0.031 0.027 −0.019 −0.036 1.000
E1 −0.064 0.039 −0.112 0.027 0.099 1.000
E2 −0.127 0.232 −0.434 0.029 0.082 0.871 1.000
E3 0.009 0.020 −0.020 −0.016 −0.029 0.101 0.080 1.000
−ΔHC3H8 0.064 −0.028 0.103 −0.032 −0.101 −0.994 −0.888 −0.098 1.000
−ΔHC3H6 0.049 −0.047 0.119 −0.018 −0.497 −0.285 −0.291 −0.328 0.291 1.000
(b) 7% VOx/γ-Al2O3 Catalyst Sample
ko1 1.000
ko2 0.696 1.000
ko3 0.344 −0.405 1.000
KoC3H8 −0.949 −0.881 −0.051 1.000
KoC3H6 −0.062 −0.093 0.014 0.083 1.000
E1 −0.521 −0.330 −0.169 0.449 0.096 1.000
E2 −0.670 −0.255 −0.507 0.535 0.063 0.705 1.000
E3 −0.043 0.016 −0.067 0.024 −0.103 −0.032 −0.019 1.000
−ΔHC3H8 0.625 0.474 0.143 −0.588 −0.104 −0.928 −0.823 0.039 1.000
−ΔHC3H6 0.129 0.121 0.017 −0.136 −0.477 −0.066 −0.076 −0.329 0.086 1.000
(c) 10% VOx/γ-Al2O3 Catalyst Sample
ko1 1.000
ko2 0.838 1.000
ko3 0.381 −0.159 1.000
KoC3H8 −0.965 −0.949 −0.143 1.000
KoC3H6 −0.122 −0.118 −0.070 0.127 1.000
E1 0.696 0.588 0.290 −0.678 −0.096 1.000
E2 0.591 0.644 −0.044 −0.639 −0.074 0.834 1.000
E3 0.091 0.090 0.032 −0.094 −0.503 0.046 0.017 1.000
−ΔHC3H8 −0.648 −0.516 −0.293 0.612 0.083 −0.961 −0.892 −0.030 1.000
−ΔHC3H6 0.002 −0.023 0.060 0.008 −0.382 0.040 0.002 −0.196 −0.033 1.000

Table 6. Reported Activation Energies for Main Products Formation in Ethane ODH
activation energy of formation (kJ/mol)
catalyst C3H6 carbon oxides reference
5% VOx/γ-Al2O3 124.9 (COx) 52.8a (COx) 52.53b (present study)
7% VOx/γ-Al2O3 115.1 (COx) 51.1a (COx) 52.7b (present study)
10% VOx/γ-Al2O3 109.5 (COx) 45.6a (COx) 53.7b (present study)
10% VOx/γ-Al2O3 82 (CO) 35b (CO2) 52.2b 81
2.1%VOx/γ-Al2O3 113.6 (CO2) 87.5b 82
7.5% V/γ-Al2O3 100 (CO) 50b (CO2) 46b 46
10% V/γ-Al2O3 98 (CO) 55b (CO2) 49b 46
12.5% V/γ-Al2O3 92 (CO) 83b (CO2) 75b 46
24%V/MgO 114 (CO) 72a (CO) 101b (CO2) 42a (CO2) 92b 43
0.6%VOx/SiO2 147 (CO2) 96b 82
a
Formation from C3H8. bFormation from C3H6.

2. The selectivity for propylene increased and that for COx oxygen-free conditions, the oxygen from the catalyst
decreased as the degree of reduction of the catalyst lattice is consumed by ODH reactions.
increased with propane injections. This shows that a 3. A parallel-series “triangular” reaction network was
controlled degree of catalyst reduction is needed in order proposed for propane ODH over VOx/γ-Al2O3 catalysts
to obtain higher propylene selectivity. Under such where there is a propylene formation step, a competitive

M dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

direct propane combustion step, and propylene XC3H8 = propane conversion, (%)
combustion step. Si = selectivity of component i (%)
4. This heterogeneous kinetic model based on a Marsvan Yi = yield of product i
Krevelen mechanism, allowed us to describe the effect of SBET = Brunauer−Emmet−Teller specific surface area (cm2/
vanadia loading on propane ODH reaction over VOx/γ- g)
Al2O3 catalysts. The proposed kinetic model accounts for Nopropane = Initial moles of propane, (mols)
all reactants (propane) and products (propylene and Wcat = weight of catalyst, (g)
COx). On this basis, rate equations are developed Ei = activation energy of desorption, (kJ/mol)
including both reactant adsorption and catalytic reaction −ΔHi = heat of adsorption, (kJ/mol)
on the catalyst surface. koi = pre-exponential factor (mol gcat−1 s−1)
5. The change in degree of oxidation of the catalyst during Kd = adsorption constant (cm3/mol)
oxygen-free ODH reactions was effectively modeled and Tm = centering temperature (K)
included in the rate equations by using a decay function R = the universal gas constant
based on reactant conversion. The proposed kinetic s = surface-adsorbed species
model was established with a DOF of 950. This model g = gas phase species
satisfactorily predicts the ODH reaction of propane Greek Symbols
under the selected reaction conditions over the various β = degree of reduction of catalysts, (%)
VOx/γ-Al2O3 catalysts with different vanadium loadings. θdes = surface coverage of adsorbed species
6. The determined kinetic parameters included the pre- ν = stoichiometric number
exponential factor, koi , the adsorption constant, Koi , the λ = catalyst deactivation constant established via numerical
activation energies, Ei, for the propene formation, for every injection in the consecutive injection series using
propane oxidation, and propene oxidation and desorp- propane conversion at 5 s, 10 s, 15 s, and 20 s
tion energies, −ΔHi. ϕ = degree of oxidation of the catalyst (% oxidation)
7. The intrinsic kinetic parameters were estimated using a σ = standard deviation
nonlinear least-squares fit regression. The calculated Abbreviations
kinetic parameters were able to predict the observed ODH = oxidative dehydrogenation
outlet mass fractions of all carbon containing compounds MW = molecular weight
and were found to vary with vanadia loading. CREC = Chemical Reactor Engineering Center
8. The calculated pre-exponential factors (ko1, ko2, and ko3) XRD = X-ray diffraction
increased as the vanadium loading was augmented. In TPR = Temperature-programmed reduction
addition, the activation energies for COx formation (E2 TPO = temperature-programmed oxidation
and E3) were consistently smaller than the one for TPD = temperature-programmed desorption
propylene formation (E1).


LRS = laser Raman spectroscopy
XPS = X-ray photoelectron spectroscopy
ASSOCIATED CONTENT FID = flame ionization detector
* Supporting Information
S TCD = thermal conductivity detector
Experimental results of propane ODH in the CREC riser AOS = average oxidation state
simulator using the various VOx/γ-Al2O3 catalyst samples for
10 consecutive propane injections (pulse) without catalyst
regeneration between injections; “λ” deactivations constants
■ REFERENCES
(1) Blasco, T.; López-Nieto, J. M. Oxidative Dehydrogenation of
obtained via numerical regression for every propane injection Short Chain Alkanes on Supported Vanadium Oxide Catalysts. Appl.
and data at 5 s, 10 s, 15 s, and 20 s. This material is available Catal. A Gen. 1997, 157, 117.
free of charge via the Internet at http://pubs.acs.org. (2) Mamedov, E. A.; Corberan, V. C. Oxidative Dehydrogenation of


Lower Alkanes on Vanadium Oxide-Based Catalysts. The Present State
AUTHOR INFORMATION of the Art and Outlooks. Appl. Catal. A Gen. 1995, 127, 1.
(3) Baerns, M.; Buyevskaya, O.; Mulla, S. A. R. A Comparative Study
Corresponding Author on Non-Catalytic and Catalytic Oxidative Dehydrogenation of Ethane
*Tel: +1-519-661-2144. Fax: +1-519-850-2931. E-mail: to Ethylene. Appl. Catal. A Gen. 2002, 226, 73.
hdelasa@eng.uwo.ca. (4) Shen, Z.; Liu, J.; Xu, H.; Yue, Y.; Hua, W.; Shen, W.
Notes Dehydrogenation of Ethane to Ethylene over a Highly Efficient
The authors declare no competing financial interest. Ga2O3/HZSM-5 Catalyst in the Presence of CO2. Appl. Catal. A Gen.


2009, 356, 148.
ACKNOWLEDGMENTS (5) Č apek, L.; Bulánek, R.; Adam, J.; Smoláková, L.; Sheng-Yang, H.;
Č ičmanec, P. Oxidative Dehydrogenation of Ethane over Vanadium-
The authors wish to acknowledge the National Sciences and Based Hexagonal Mesoporous Silica Catalysts. Catal. Today 2009, 141,
Engineering Research Council of Canada (NSERC) for their 282.
financial support of this research and the Saudi Arabian Oil (6) Shi, X.; Ji, S.; Wang, K. Oxidative Dehydrogenation of Ethane to
Company (Saudi ARAMCO), Saudi Arabia, for the scholarship Ethylene with Carbon Dioxide over Cr−Ce/SBA-15 Catalysts. Catal.
awarded for the development of this Ph.D. research at the Lett. 2008, 125, 331.
University of Western Ontario. (7) Č apek, L.; Adam, J.; Grygar, T.; Bulánek, R.; Vradman, L.;


Košová-Kučerová, G.; Č ičmanec, P.; Knotek, P. Oxidative Dehydro-
genation of Ethane over Vanadium Supported on Mesoporous
NOTATIONS Materials of M41S Family. Appl. Catal. A Gen. 2008, 342, 99.
ri = Rate of reaction i (mol/g.s) (8) Klose, F.; Wolff, T.; Lorenz, H.; Seidelmorgenstern, A.;
VOx = vanadium oxide surface species Suchorski, Y.; Piorkowska, M.; Weiss, H. Active Species on Γ-

N dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

Alumina-Supported Vanadia Catalysts: Nature and Reducibility. J. (31) Bell, A. T.; Khodakov, A.; Olthof, B.; Iglesia, E. Structure and
Catal. 2007, 247, 176. Catalytic Properties of Supported Vanadium Oxides: Support Effects
(9) Danica, B.; Desislava, A.; Mirko, P.; Stefan, H. An Experimental on Oxidative Dehydrogenation Reactions. J. Catal. 1999, 181, 205.
Study of the Partial Oxidation of Ethane to Ethylene in a Shallow (32) Chen, K.; Xie, S.; Bell, A. T.; Iglesia, E. Structure and Properties
Fluidized Bed Reactor. J. Serbian Chem. Soc. 2007, 72, 183. of Oxidative Dehydrogenation Catalysts Based on MoO3/Al2O3. J.
(10) Haddad, N.; Bordes Richard, E.; Hilaire, L.; Barama, A. Catal. 2001, 198, 232.
Oxidative Dehydrogenation of Ethane to Ethene on Alumina- (33) Grabowski, R.; Sloczynski, J. Kinetics of Oxidative Dehydrogen-
Supported Molybdenum-Based Catalysts Modified by Vanadium and ation of Propane and Ethane on VOx/SiO2 Pure and with Potassium
Phosphorus. Catal. Today 2007, 126, 256. Additive. Chem. Eng. Process. 2005, 44, 1082.
(11) Karamullaoglu, G.; Dogu, T. Oxidative Dehydrogenation of (34) Grabowski, R. Kinetics of Oxidative Dehydrogenation of C2-C3
Ethane over Chromium-Vanadium Mixed Oxide and Chromium Alkanes on Oxide Catalysts. Catal. Rev. 2006, 48, 199.
Oxide Catalysts. Ind. Eng. Chem. Res. 2007, 46, 7079. (35) Lemonidou, A. A. Oxidative Dehydrogenation of C4 Hydro-
(12) Tope, B.; Zhu, Y.; Lercher, J. A. Oxidative Dehydrogenation of carbons over VMgO Catalyst  Kinetic Investigations. Appl. Catal. A
Ethane over Dy2O3/MgO Supported LiCl Containing Eutectic Gen. 2001, 216, 277.
Chloride Catalysts. Catal. Today 2007, 123, 113. (36) Machli, M.; Boudouris, C.; Gaab, S.; Find, J.; Lemonidou, A. A.;
(13) Cavani, F.; Ballarini, N.; Cericola, A. Oxidative Dehydrogen- Lercher, J. A. Kinetic Modelling of the Gas Phase Ethane and Propane
ation of Ethane and Propane: How far from Commercial Oxidative Dehydrogenation. Catal. Today 2006, 112, 53.
Implementation? Catal. Today 2007, 127, 113. (37) Klose, F.; Joshi, M.; Hamel, C.; Seidelmorgenstern, A. Selective
(14) Lemonidou, A. A.; Heracleous, E. Reaction Pathways of Ethane Oxidation of Ethane over a VOx/γ-Al2O3 CatalystInvestigation of
Oxidative and Non-Oxidative Dehydrogenation on Γ-Al2O3 Studied the Reaction Network. Appl. Catal. A Gen. 2004, 260, 101.
by Temperature-Programmed Reaction (TP-Reaction). Catal. Today (38) Rahman, F.; Loughlin, K. F.; Al-Saleh, M. A.; Saeed, M. R.;
2006, 112, 23. Tukur, N. M.; Hossain, M. M.; Karim, K.; Mamedov, E. A. Kinetics
(15) Solsona, B.; Dejoz, A.; Garcia, T.; Concepcion, P.; López-Nieto, and Mechanism of Partial Oxidation of Ethane to Ethylene and Acetic
J. M.; Vazquez, M.; Navarro, M. T. Molybdenum−Vanadium Acid over MoV Type Catalysts. Appl. Catal. A Gen. 2010, 375, 17.
Supported on Mesoporous Alumina Catalysts for the Oxidative (39) Chen, K.; Iglesia, E.; Bell, A. T. Isotopic Tracer Studies of
Dehydrogenation of Ethane. Catal. Today 2006, 117, 228. Reaction Pathways for Propane Oxidative Dehydrogenation on
(16) Heracleous, E.; Lemonidou, A. A. Ni−Nb−O Mixed Oxides as Molybdenum Oxide Catalysts. J. Phys. Chem. B 2001, 105, 646.
Highly Active and Selective Catalysts for Ethene Production via (40) Bell, A. T.; Iglesia, E.; Argyle, M. D.; Chen, K. Ethane Oxidative
Ethane Oxidative Dehydrogenation. Part II: Mechanistic Aspects and Dehydrogenation Pathways on Vanadium Oxide Catalysts. J. Phys.
Kinetic Modeling. J. Catal. 2006, 237, 175. Chem. B 2002, 106, 5421.
(17) Heracleous, E.; Lemonidou, A. A. Ni−Nb−O Mixed Oxides as (41) Bell, A. T.; Dinse, A.; Schomäcker, R. The Role of Lattice
Highly Active and Selective Catalysts for Ethene Production via Oxygen in the Oxidative Dehydrogenation of Ethane on Alumina-
Ethane Oxidative Dehydrogenation. Part I: Characterization and Supported Vanadium Oxide. Phys. Chem. Chem. Phys. 2009, 11, 6119.
Catalytic Performance. J. Catal. 2006, 237, 162. (42) Lemonidou, A. A.; Heracleous, E.; Lercher, J. A. Mechanistic
(18) Botella, P.; Dejoz, A.; López-Nieto, J. M.; Concepcion, P.; Features of the Ethane Oxidative Dehydrogenation by in Situ FTIR
Vazquez, M. Selective Oxidative Dehydrogenation of Ethane over Spectroscopy over a MoO3/Al2O3 Catalyst. Appl. Catal. A Gen. 2004,
MoVSbO Mixed Oxide Catalysts. Appl. Catal. A Gen. 2006, 298, 16. 264, 73.
(19) Nakagawa, K.; Miyake, T.; Konishi, T.; Suzuki, T. Oxidative (43) Ramos, R.; Pinal, M. P.; Menendez, M.; Santamaria, J. Patience,
Dehydrogenation of Ethane to Ethylene over NiO Loaded on High G. S. Oxidative Dehydrogenation of Propane to Propene, 1: Kinetic
Surface Area MgO. J. Mol. Catal. A Chem. 2006, 260, 144. Study on V/MgO. Can. J. Chem. Eng. 2001, 79, 891.
(20) López-Nieto, J. M.; Botella, P.; García-González, E.; Dejoz, A.; (44) Creaser, D.; Andersson, B. Oxidative Dehydrogenation of
Vazquez, M.; González-Calbet, J. Selective Oxidative Dehydrogenation Propane over V-Mg-O: Kinetic Investigation by Nonlinear Regression
of Ethane on MoVTeNbO Mixed Metal Oxide Catalysts. J. Catal. Analysis. Appl. Catal. A Gen. 1996, 141, 131.
2004, 225, 428. (45) Creaser, D.; Andersson, B.; Hudgins, R. R.; Silverston, P. L.
(21) Anon, R. New Propylene Production Technologies Hold Great Transient Kinetic Analysis of the Oxidative Dehydrogenation of
Promise. Oil Gas J. 2004, 102, 50. Propane. J. Catal. 1999, 182, 264.
(22) Bhasin, M. Is True Ethane Oxydehydrogenation Feasible? Top. (46) Rao, T. V. M; Deo, G. Kinetic Parameter Analysis for Propane
Catal. 2003, 23, 145. ODH: V2O5/Al2O3 and MoO3/Al2O3 Catalysts. AIChE J. 2007, 53,
(23) Resasco, D.; Haller, G. Catalytic Dehydrogenation of Lower 1538.
Alkanes. Catal. R. Soc. Chem. 1994, 11. (47) Chen, K.; Iglesia, E.; Bell, A. T. Kinetic Isotopic Effects in
(24) Chan, K. Y. G.; Inal, F.; Senkan, S. Suppression of Coke Oxidative Dehydrogenation of Propane on Vanadium Oxide Catalysts.
Formation in the Steam Cracking of Alkanes: Ethane and Propane. J. Catal. 2000, 192, 197.
Ind. Eng. Chem. Res. 1998, 37, 901. (48) Creaser, D.; Andersson, B. Oxidative Dehydrogenation of
(25) Dharia, D.; Letzsch, W.; Kim, H.; McCue, D.; Chapin, L. Propane over V-Mg-O: Kinetic Investigation by Nonlinear Regression
Increase Light Olefins Production. Hydrocarb. Process. 2004, 83, 61. Analysis. Appl. Catal. A Gen. 1996, 141, 131.
(26) Ren, T.; Patel, M.; Blok, K. Olefins from Conventional and (49) Dinse, A.; Khennache, S.; Frank, B.; Hess, C.; Herbert, R.;
Heavy Feedstocks: Energy Use in Steam Cracking and Alternative Wrabetz, S.; Schlögl, R.; Schomäcker, R. Oxidative Dehydrogenation
Processes. Energy 2006, 31, 425. of Propane on Silica (SBA-15) Supported Vanadia Catalysts: A Kinetic
(27) Farrauto, R. J.; Bartholomew, C. H. Fundamentals of Industrial Investigation. J. Mol. Catal. A Chem. 2009, 307, 43.
Catalytic Processes; 1st ed.; Chapman & Hall: London, 1997. (50) Balcaen, V.; Sack, I.; Olea, M.; Marin, G. B. Transient Kinetic
(28) Chen, K.; Xie, S.; Bell, A. T.; Iglesia, E. Alkali Effects on Modeling of the Oxidative Dehydrogenation of Propane over a
Molybdenum Oxide Catalysts for the Oxidative Dehydrogenation of Vanadia-Based Catalyst in the Absence of O2. Appl. Catal. A Gen.
Propane. J. Catal. 2000, 195, 244. 2009, 371, 31.
(29) Kung, H. Oxidative Dehydrogenation of Light (C2 to C4) (51) Bottino, A.; Capannelli, G.; Comite, A.; Storace, S.; Felice, R. D.
Alkanes. Adv. Catal. 1994, 40, 1. Kinetic Investigations on the Oxidehydrogenation of Propane over
(30) Chen, K.; Khodakov, A.; Yang, J.; Bell, A. T.; Iglesia, E. Isotopic Vanadium Supported on Γ-Al2O3. Chem. Eng. J. 2003, 94, 11.
Tracer and Kinetic Studies of Oxidative Dehydrogenation Pathways on (52) Grabowski, R.; Pietrzyk, S.; Słoczy, J.; Genser, F.; Wcisło, K.;
Vanadium Oxide Catalysts. J. Catal. 1999, 333, 325. Swierkosz, B. G.-. Kinetics of the Propane Oxidative Dehydrogenation

O dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX


Industrial & Engineering Chemistry Research Article

on Vanadia/Titania Catalysts from Steady-State and Transient (74) Hossain, M. M.; Atanda, L.; Al-Yassir, N.; Al-Khattaf, S. Kinetics
Experiments. Appl. Catal. A Gen. 2002, 232, 277. Modeling of Ethylbenzene Dehydrogenation to Styrene over a
(53) Leveles, L. Oxidative Conversion of Propane over Lithium- Mesoporous Alumina Supported Iron Catalyst. Chem. Eng. J. 2012,
Promoted Magnesia Catalyst I. Kinetics and Mechanism. J. Catal. 207−208, 308.
2003, 218, 296. (75) Al-Ghamdi, S. A.; Volpe, M.; Hossain, M. M.; de Lasa, H. I.
(54) Kumar, C. P.; Gaab, S.; Müller, T. E.; Lercher, J. A. Oxidative VOx/c-Al2O3 Catalyst for Oxidative Dehydrogenation of Ethane to
Dehydrogenation of Light Alkanes on Supported Molten Alkali Metal Ethylene: Desorption Kinetics and Catalytic Activity. Appl. Catal. A
Chloride Catalysts. Top. Catal. 2008, 50, 156. Gen. 2013, 450, 120.
(55) Al-Zahrani, S. M.; Abasaeed, A. E.; Putra, M. Kinetics of (76) Ginsburg, J.; de Lasa, H. I. Catalytic Gasification of Biomass in
Oxidehydrogenation of Propane over Alumina-Supported Sr−V−Mo CREC Fluidized Riser Simulator. Int. J. Chem. React. Eng. 2005, 3, A38.
Catalysts. Catal. Commun. 2012, 26, 98. (77) Pekediz, A.; Kraemer, D. W.; Chabot, J.; de Lasa, H. I. Mixing
(56) Late, L.; Blekkan, E. A. Kinetics of the Oxidative Dehydrogen- Patterns in a Novel Riser Simulator. Appl. Sci. 1992, 133.
ation of Propane over a VMgO Catalyst. J. Nat. Gas Chem. 2002, 11, (78) Hossain, M. M.; de Lasa, H. I. Reduction and Oxidation Kinetics
33. of Co−Ni/Al2O3 Oxygen Carrier Involved in a Chemical-Looping
(57) Gascón, J.; Valenciano, R.; Téllez, C.; Herguido, J.; Menendez, Combustion Cycles. Chem. Eng. Sci. 2010, 65, 98.
M. A Generalized Kinetic Model for the Partial Oxidation of N-Butane (79) Jarosch, K.; El Solh, T.; de Lasa, H. I. Modelling the Catalytic
to Maleic Anhydride under Aerobic and Anaerobic Conditions. Chem. Steam Reforming of Methane: Discrimination between Kinetic
Eng. Sci. 2006, 61, 6385. Expressions Using Sequentially Designed Experiments. Chem. Eng.
(58) López-Nieto, J. M.; Soler, J.; Concepcion, P.; Herguido, J.; Sci. 2002, 57, 3439.
(80) Moreira, J.; Serrano, B.; Ortiz, A.; de Lasa, H. I. A Unified
Menendez, M.; Santamaria, J. Oxidative Dehydrogenation of Alkanes
Kinetic Model for Phenol Photocatalytic Degradation over TiO2
over V-Based Catalysts: Influence of Redox Properties on Catalytic
Photocatalysts. Chem. Eng. Sci. 2012, 78, 186.
Performance. J. Catal. 1999, 332, 324.
(81) Routray, K.; Reddy, K. R. S.; Deo, G. Oxidative Dehydrogen-
(59) Sadykov, V. A.; Pavlova, S.; Saputina, N.; Zolotarskii, I.;
ation of Propane on V2O5/Al2O3 and V2O5/TiO2 Catalysts:
Pakhomov, N.; Moroz, E.; Kuzmin, V.; Kalinkin, A. Oxidative Understanding the Effect of Support by Parameter Estimation. Appl.
Dehydrogenation of Propane over Monoliths at Short Contact Catal. A Gen. 2004, 265, 103.
Times. Catal. Today 2000, 61, 93. (82) Dinse, A.; Frank, B.; Hess, C.; Habel, D.; Schomäcker, R.
(60) Leveles, L. Oxidative Conversion of Propane over Lithium- Oxidative Dehydrogenation of Propane over Low-Loaded Vanadia
Promoted Magnesia Catalyst II. Active Site Characterization and Catalysts: Impact of the Support Material on Kinetics and Selectivity. J.
Hydrocarbon Activation. J. Catal. 2003, 218, 307. Mol. Catal. A Chem. 2008, 289, 28.
(61) Zhao, R.; Wang, J.; Dong, Q.; Liu, J. Selective Oxidation of
Propane by Lattice Oxygen of Vanadium−Phosphorous Oxide in a
Pulse Reactor. J. Nat. Gas Chem. 2005, 14, 88.
(62) Fukudome, K.; Ikenaga, N.; Miyake, T.; Suzuki, T. Oxidative
Dehydrogenation of Propane Using Lattice Oxygen of Vanadium
Oxides on Silica. Catal. Sci. Technol. 2011, 1, 987.
(63) Yoon, Y. S.; Ueda, W.; Moro-oka, Y. Oxidative Dehydrogen-
ation of Propane over Magnesium Molybdate Catalysts. Catal. Lett.
1995, 35, 57.
(64) Creaser, D.; Andersson, B.; Hudgins, R. R.; Silverston, P. L.
Transient Study of Oxidative Dehydrogenation of Propane. Appl.
Catal. A Gen. 1999, 187, 147.
(65) Stern, D. L.; Grasselli, R. K. Propane Oxydehydrogenation over
Molybdate-Based Catalysts. J. Catal. 1997, 167, 550.
(66) Rubio, O.; Herguido, J.; Menendez, M. Oxidative Dehydrogen-
ation of N-Butane on V/MgO Catalysts-Kinetic Study in Anaerobic
Conditions. Chem. Eng. Sci. 2003, 58, 4619.
(67) Soler, J.; López-Nieto, J. M.; Herguido, J.; Menendez, M.;
Santamaria, J. Oxidative Dehydrogenation of N-Butane in a Two-Zone
Fluidized-Bed Reactor. Ind. Eng. Chem. Res. 1999, 38, 90.
(68) Rubio, O.; Herguido, J.; Menendez, M. Oxidative Dehydrogen-
ation of Butane in an Interconnected Fluidized-Bed Reactor. AIChE J.
2004, 50, 1510.
(69) Al-Ghamdi, S. A.; Hossain, M. M.; Lasa, H. I. De. Kinetic
Modeling of Ethane Oxidative Dehydrogenation over VOX/Al2O3
Catalyst in a Fluidized-Bed Riser Simulator. Ind. Eng. Chem. Res.
2013, 52, 5235−5244.
(70) Al-Ghamdi, S. A.; Volpe, M.; Hossain, M. M.; de Lasa, H. I.
VOx/c-Al2O3 Catalyst for Oxidative Dehydrogenation of Ethane to
Ethylene: Desorption Kinetics and Catalytic Activity. Appl. Catal. A
Gen. 2013, 450, 120.
(71) Al-Ghamdi, S. A.; de Lasa, H. I. Propane Oxidative
Dehydrogenation over a VOx/Al2O3 Catalyst: Characterization and
Catalytic Performance. Fuel 2013, submitted.
(72) De Lasa, H. I. Canadian Patent 1,284,017, (1991). USA Pat.
5,102,628, 1992.
(73) De Lasa, H. I.; Al-Khattaf, S. Catalytic Cracking of Cumene in a
Riser Simulator: A Catalyst Activity Decay Model. Ind. Eng. Chem. Res.
2001, 5398.

P dx.doi.org/10.1021/ie404064j | Ind. Eng. Chem. Res. XXXX, XXX, XXX−XXX

You might also like