You are on page 1of 61

Accepted Manuscript

Modeling of oxidative dehydrogenation of ethane to ethylene on a MoVTeNbO/


TiO2 catalyst in an industrial-scale packed bed catalytic reactor

Gamaliel Che-Galicia, Richard S. Ruiz-Martínez, Felipe López-Isunza, Carlos


O. Castillo-Araiza

PII: S1385-8947(15)00838-4
DOI: http://dx.doi.org/10.1016/j.cej.2015.05.128
Reference: CEJ 13770

To appear in: Chemical Engineering Journal

Received Date: 5 March 2015


Revised Date: 20 May 2015
Accepted Date: 22 May 2015

Please cite this article as: G. Che-Galicia, R.S. Ruiz-Martínez, F. López-Isunza, C.O. Castillo-Araiza, Modeling of
oxidative dehydrogenation of ethane to ethylene on a MoVTeNbO/TiO2 catalyst in an industrial-scale packed bed
catalytic reactor, Chemical Engineering Journal (2015), doi: http://dx.doi.org/10.1016/j.cej.2015.05.128

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Modeling of oxidative dehydrogenation of ethane to ethylene on

a MoVTeNbO/TiO2 catalyst in an industrial-scale packed bed

catalytic reactor

Gamaliel Che-Galicia, Richard S. Ruiz-Martínez, Felipe López-Isunza, Carlos O. Castillo-

Araiza*

Grupo de Procesos de Transporte y Reacción en Sistemas Multifásicos, Depto. de Ingeniería de

Procesos e Hidráulica, Universidad Autónoma Metropolitana - Iztapalapa, Av. San Rafael Atlixco

No. 186, C.P. 09340, México D.F., Mexico.

*Corresponding author

e-mail addresses: coca@xanum.uam.mx (C.O. Castillo-Araiza) and

gamaliel.che@gmail.com (G. Che-Galicia)

Tel: + 52 55 5804 4648

1
Abstract

The oxidative dehydrogenation of ethane (ODH-Et) seems the most promising alternative

to produce ethylene compared to conventional processes. Nevertheless, there is not even a

pilot plant for ODH-Et worldwide nowadays. This work presents the simulation of the

catalytic behavior of a highly active and selective MoVTeNbO catalyst for the ODH-Et in a

wall-cooled industrial-scale packed-bed reactor with a tube to particle diameter ratio equal

to 3.12. The feasibility of using this complex yet necessary reactor design, as well as the

influence of operating conditions on conversion and yield along the reactor are analyzed.

The simulations are carried out using a two-dimensional pseudo-heterogeneous model,

which makes use of both a reliable kinetic model and reliable transport parameters.

Specifically, the kinetic model, obtained from lab-scale experimental data, is coupled to the

reactor model accounting for transport phenomena wherein the effect of hydrodynamics on

heat transfer is assessed from independent experiments in absence of reaction in an

industrial-scale reactor. The developed kinetics successfully accounts for both the

inhibiting effect of adsorbed water on oxidative dehydrogenation and total oxidations and

the effect of the inlet partial pressure of oxygen and ethane on conversion and selectivity.

Besides, the reactor model elucidates the importance of accounting for the role of the

hydrodynamics on the heat transport in order to have reliable conversion and yield

predictions. From the parametric sensitivity study, the temperature of the coolant fluid and

the inlet ethane concentration are variables for tuning the reactor performance.

Key words: Oxidative dehydrogenation reaction; low dt/dp packed bed reactor;

Hydrodynamics; Ethylene; Reaction kinetics; MoVTeNbO catalyst.

2
Abbreviations: BIC, Bayesian information criterion; BLA, boundary layer approximation;

CA, classical approach; dt/dp, tube-to-particle diameter ratio; ER, Eley-Rideal; GC, gas

chromatograph; HTP, heat transport parameters; LHHW, Langmuir-Hinshelwood-Hougen-

Watson; MvK, Mars-van Krevelen; ODEs, ordinary differential equations; ODH-Et,

oxidative dehydrogenation of ethane; PL, power law.

3
1 Introduction

Ethylene is one of the principal building blocks in petrochemical industry as it is used as

intermediate for the production of many valuable products, namely polyethylene, ethylene

oxide, ethylene dichloride and ethylene benzene, among other important chemical

compounds [1,2]. Industry and academy have focused their research on finding competitive

technologies to produce ethylene [2-5] so as to face its increasing demand and also, as an

effort of reducing the environmental impact and energy consumption of the conventional

commercial processes for producing this olefin, i.e., pyrolysis, fluid catalytic cracking and

catalytic dehydrogenation [5,6]. The oxidative dehydrogenation of ethane (ODH-Et), in this

regard, seems to be the most promising alternative since it offers many advantages

compared to conventional processes, for instance: it overcomes equilibrium limitations;

permits lower temperature operation (<500ºC) and, hence, energy saving; it decreases the

number of side reactions; and, also, avoids catalyst deactivation by coke formation due to

the presence of molecular oxygen contained in the reaction mixture [4,5,7,8].

Notwithstanding these advantages there are two industrial challenges for producing

ethylene via ODH-Et. One of them is related to the design of an active and selective

catalyst to produce ethylene, while the other one has to do with the design of the reactor

technology. In fact today, to the best of our knowledge, there is not even a pilot plant for

ODH-Et worldwide.

A broad variety of catalytic formulations to produce ethylene out of ethane in

presence of oxygen have been investigated [9-19]. From these formulations, the

multimetallic mixed oxide catalyst based on Mo, V, Te and Nb is, undoubtedly, the most

efficient and effective material since it presents an outstanding activity and selectivity to

4
ethylene, apparently attributed to the presence of the M1 crystalline phase. Particularly, V

species are the active sites in ethane activation, Mo species enhance the catalytic activity of

V atoms, Te species are associated with the formation of the active and selective crystalline

phase M1, and Nb species leads to enhance the selectivity to ethylene [11,13,20-23]. From

this end, envisaging a possible future application of this material in ODH-Et to a

commercial scale requires, as a first step, of the conceptual design of the catalytic reactor

technology. Nevertheless, owing to the exothermicity of the ODH-Et reaction, mainly

related to the unavoidably heat released during the formation of COX (CO and CO2) out of

ethane and ethylene, the reactor design is as yet a complex chemical engineering

assignment, as it is analyzed in the following.

The wall-cooled catalytic reactor with low tube-to-particle diameter ratio (dt/dp < 5)

seems a suitable option to commercially achieve the ODH-Et over the MoVTeNbO

formulation [24,25], since this configuration and catalyst design promote the fast removal

of the heat generated, maximize the desirable product yield, and avoids strong catalyst

deactivation by a sintering process [26-34], as observed for other well-known studied

exothermic partial oxidation reactions. In fact, several researches [35-41] have considered

the wall-cooled packed bed catalytic reactor for simulating the possible performance of the

ODH-Et over catalytic systems different to the MoVTeNbO formulation. Nevertheless,

these studies are questionable since they do not make use of reliable kinetic models and

neglect the effect of the low dt/dp on the void fraction profile and, hence, on velocity field

and its role on the heat transport [42-46].

The conceptual design of the industrial reactor for the ODH-Et over a MoVTeNbO

catalyst requires of both a reliable kinetic model and reliable transport parameters. A few

5
number of kinetic models have been developed to describe the macroscopic mechanism of

ODH-Et, namely Power Law (PL), Langmuir-Hinshelwood-Hougen-Watson (LHHW),

Eley Rideal (ER), Mars-van Krevelen (MvK) or a combination of MvK and LHHW (MvK–

LHHW) [47-58], nevertheless, these models are specific for every catalyst formulation.

Specifically for the MoVTeNbO catalyst, the LHHW model, developed by our research

group, is one of the most suitable one nowadays [57] since other kinetic models present a

lack of statistical and phenomenological significance. Notwithstanding LHHW is a

potential kinetic model to overcome the conceptual design of the industrial ODH-Et reactor

using the MoVTeNbO catalytic system, some aspects need to be accounted for, i.e., the

relatively weak effect of oxygen partial pressure on reaction rates, the high affinity of water

to adsorb on active sites affecting negatively reaction rates and the significative effect of

inlet partial pressure of ethane on reaction rates.

The aim of this study is to model the industrial performance of a highly

active/selective MoVTeNbO catalyst for the production of ethylene out of ethane via ODH-

Et in a wall-cooled packed bed catalytic reactor presenting a dt/dp equal to 3.12. This

industrial-scale reaction system is simulated using a two-dimensional pseudo-

heterogeneous model coupled to a kinetic model developed herein for MoVTeNbO

formulation and accounts, in some detail, for the effect of hydrodynamics on heat transfer

through the proper estimation of heat transport parameters from pilot plant experiments in

absence of reaction.

6
2 Reaction Kinetics

2.1 Catalyst

A MoVTeNbO catalyst with a nominal atomic ratio of Mo:V:Te:Nb equal to

1:0.24:0.24:0.18 prepared by slurry method is used. From XRD pattern using Rietveld

refinement, the crystalline phase composition (wt %) is analyzed to be the M1 major

crystalline phase present in the catalyst (92 wt % M1 and 8 wt % M2 crystalline phase)

[18]. A specific surface area of 7 m2 gcat-1 is obtained from N2 physisorption in an

AUTOSORB-I instrument. Details concerning the preparation as well as the

physicochemical characteristics of this catalyst are given elsewhere [18].

2.2 Experimental Data

The data used in the kinetic modeling is obtained from experiments carried out in a

laboratory isothermal fixed-bed reactor at atmospheric pressure. The fixed-bed reactor,

made of quartz, has an internal diameter of 1.0×10-2 m and length of 4.0×10-2 m.

Experiments were performed using a bed consisting of 0.60 g of catalyst with average

particle size of 150 µm. The reaction feedstock was composed of a mixture of ethane (or

ethylene), oxygen and nitrogen with a varying composition depending on the experiment.

The experimental conditions and conversion ranges are listed in Table 1. The reactor

effluent was analyzed online periodically by gas chromatography (GC). A detailed

description of the equipment has been reported in a previous study [57].

Table 1 is shown here.

7
2.3 Reaction rates

The reaction network for ODH-Et over MoVTeNbO catalysts is proposed taking into

account the experimental observations given in [57]. Both parallel and consecutive

reactions are considered, in which the species detected by GC (ethane, ethylene, carbon

monoxide and carbon dioxide) are only included. In this regard:

Ethylene is produced from oxidative dehydrogenation of ethane (r1):

C2H6 + 0.5O2 → C2H4 + H2O (1)

Ethane can also be oxidized to form CO2 (r2) and/or CO (r3):

C2H6 + 3.5O2 → 2CO2 + 3H2O (2)

C2H6 + 2.5O2 → 2CO + 3H2O (3)

Ethylene can also react to form CO2 (r4) and/or CO (r5).

C2H4 + 3O2 → 2CO2 + 2H2O (4)

C2H4 + 2O2 → 2CO + 2H2O (5)

Table 2 shows the reaction mechanism considered for building the kinetic model. The

mechanism accounts for some of the particular aspects of the MoVTeNbO formulation

reported in the literature [57], namely a relatively weak effect of oxygen partial pressure on

reaction rates, a high tendency of water to adsorb on active sites and the significative effect

of inlet partial pressure of ethane on conversion and selectivity. Thus, the model accounts

for the following assumptions: (i) there is a single type of active sites (S) on the catalyst

surface, (ii) the only compounds adsorbed on the active sites are oxygen and water, (iii)

oxygen adsorption is dissociative (step A) while water adsorption is associative (step B),

8
(iv) surface reaction steps (steps 1 to 5) are considered to be fast taking place over a finite

number of active sites, (v) since the construction of the model is based on the pseudo

equilibrium approach, surface reactions are the rate-determining steps and adsorption and

desorption steps are quasi-equilibrated. In Table 2, σj is the so-called Horiuti number, that

describes the times that every adsorption-desorption and/or reaction steps need to occur in

order to complete a single catalytic cycle comprising the overall reactions (eqs. (1) to (5)).

Notice that the adsorption of molecular O2 (step A) is a crucial stage for the surface

reactions to take place (steps 1 to 5).

Table 2 is shown here.

The equations that describe the rate of reactions for ODH-Et over MoVTeNbO are

given through eqs. (6) to (11), and are written as a function of the partial pressures of the

gas phase components according to the reaction mechanism. The calculation of ri involves a

rate coefficient denoted by ki, an adsorption equilibrium coefficient for the n-th component

represented by Kn, the partial pressure of the reactant n represented by pn, the fraction

coverage of oxygen denoted by θ O and the reaction order related to the partial pressure of

the gas phase denoted by mi:

r1 = k1pC2H6 θO (6)

r2 = k 2 pC2H6 θOm2 (7)

r3 = k 3p C2H6 θmO 3 (8)

9
r4 = k 4p C2H4 θOm4 (9)

r5 = k 5 p C2H4 θmO 5 (10)

Since the fraction coverage of different species on the catalytic surface is conserved

through time, the global balance of the fraction sites corresponds to Eq. (11):

θ* + θO + θH2 O = 1 (11)

The fraction coverage of a specific component is defined as the number of sites

occupied by such a species relative to the total number of available sites corresponding to

Eq. (12) for oxygen and Eq. (13) for water:

θO = (KO2 pO2 )1 2 θ* (12)

θ H 2 O = K H 2 O p H 2 O θ* (13)

where the adsorption equilibrium coefficient for the n-th component is represented by Kn,

the partial pressure of the reactant n is represented by pn, and the fraction coverage of active

sites is denoted by θ* . The fraction of free active sites is obtained by combining eqs. (12)

and (13) resulting in Eq. (14):

1
θ* = 12
(14)
1 + (K O2 pO2 ) + K H2O pH2O

The combination of specific reaction rates gives the net reaction rate of formation for

the n-th component, which is expressed by Eq. (15):

5
R n = ∑ ν n ,i ri (15)
i =1

10
where ν n ,i is the stoichiometric coefficient of the component n in the i-th reaction, vide

eqs. (1) to (5). In order to reduce the correlation between activation energies and pre-

exponential factors as well as standard adsorption enthalpies and entropies during the

parameters estimation procedure, both Arrhenius and van’t Hoff equations are used in their

reparametrized form.

 E  1 1 
ki = exp  A'i − A,i  − *   (16)
 R  T T 

 ∆So ∆Hon  1 1 
Kn = exp  n −  − *  (17)
 R R  T T 

where A 'i is the natural logarithm of the pre-exponential factor for the i-th reaction, EA,i is

the activation energy factor for the i-th reaction, T is the reaction temperature, T* is the

averaged reaction temperature, ∆ Son is the standard adsorption entropy of component n,

∆ H on is the standard adsorption enthalpy of component n and R is the universal gas

constant.

2.3.1 Laboratory reactor model

The continuity equation for component n at the steady state assumes that the flow pattern in

the laboratory reactor is of plug flow type and that there is no radial temperature or

concentration gradient as corroborated elsewhere [57]:

dFn
= Rn (18)
dWcat

with the following initial conditions:

11
Fn = Fno , for Wcat = 0 (19)

where Fn is the molar flow rate of component n, Fno is the inlet molar flow rate of

component n and Wcat is the mass of the catalyst loaded into the reactor.

2.3.2 Parameter estimation

The development of a kinetic model must account for a proper assessment of the model

parameters from both physicochemical and statistical perspectives since these evaluations

are essential for carrying out a reliable chemical reactor design. The kinetic parameters are

estimated by minimization of the following weighted objective function:

n resp n exp

RSS(β) = ∑ w n ∑ (Fk,n − Fˆk,n )2 


β1 ,β2 ,...,βn
→ min (20)
n =1 k =1

where β is an optimal parameter vector, nexp is the number of independent experiments, nresp

is the number of responses, Fk,n and F̂k,n are the n-th experimental and predicted responses

for the k-th observation, respectively; and wn is the weight factor assigned to the n-th

response.

The subroutine VODE is used to solve the corresponding set of ODEs [59]. The

initial minimization of the objective function, vide Eq. (20), in the model regression is

carried out using the Rosenbrock method [60] and then, the ODRPACK subroutines are

called for fitting the corresponding experimental data points [61]. These subroutines can

perform either weighted orthogonal distance regression or nonlinear least square problems

for explicit and implicit models using multi-response data with an implementation of the

Levenberg-Marquardt method, as documented elsewhere [62].

12
Three criteria are used to evaluate the kinetic model. Firstly, the agreement between

the calculated and experimental molar flows is compared using the so-called parity plot,

thus, a careful assessment of both individual and global statistical significance is carried

out. The individual statistical significance of the parameters is assessed by a t-test while the

global significance of the regression is assessed by the F-test. Aside, the parameter

correlation between pairs of estimated parameters is accounted for by computing the so-

called binary linear correlation coefficients (ρij). When the value of ρij is close to ±1 a

strong linear relationship between the estimated parameters i and j occurs.

Finally, kinetic parameters have to be contrasted with a series of thermodynamic

criteria in order to determine whether they are physically meaningful or not. In relation to

the Arrhenius equation parameters, the activation energy should be lower to 210 kJ mol-1

[63] since larger values indicate the presence of catalyst deactivation, for instance, by

sintering [64]; whereas for the van’t Hoff equation parameters, estimated values of standard

adsorption enthalpy and entropy need to be confronted with a set of criteria stipulated by

Boudart [65], which are, often, overlooked in literature.

Since adsorption is an exothermic process, the adsorption enthalpy has to satisfy the

inequality:

−∆H o > 0 (21)

the negative value of the standard entropy of adsorption must be higher than zero and lower

than the corresponding standard entropy of the gas phase specie ( Son,g ), namely:

0 < −∆Son < Son,g (22)

where ∆Son is the standard entropy of adsorption.

13
The entropy in fact decreases when a gaseous molecule is transferred from a three-

dimensional phase, i.e, the gas phase, to a two-dimensional phase, i.e, the catalyst surface.

The gas phase standard entropy values of ethane, oxygen, ethylene, CO, CO2 and water are

computed at 440 ºC amounting to 275, 231, 257, 223, 246 and 218 J (mol K)-1,

respectively, based on reference [57].

An additional criterion to be satisfied during the parameter estimation for the nonisothermal

situation related to adsorption reaction, relates to the change in volume that occurs when a

gaseous molecule is adsorbed over the surface of a solid. This is specifically expressed in

the following form:

41.8 < −∆So < 51.04 − 1.4∆Ho (23)

3 Industrial-scale reactor model

The performance of a multi-tubular wall-cooled packed bed catalytic reactor is, normally,

predicted by the modeling of a single tube, assuming, in a reliable manner, that all tubes

behave almost identically [66,67]. Due to this, this work simulates the performance of the

MoVTeNbO catalytic formulation during the ODH-Et in a single tube industrial-scale

packed bed reactor operating in a non-isothermal and non-adiabatic mode. The reliability of

the simulations is based on the modeling approach proposed by Castillo-Araiza and López-

Isunza [33,34,45], which allows the prediction of the observed behavior of a wall-cooled

packed bed catalytic reactor with low dt/dp. This methodology accounts for the role of void

fraction and hydrodynamics on heat transport in the reactor through the proper

determination, using experiments in absence of reaction, of the effective heat transport

14
parameters (HTP) such as the radial and axial effective thermal conductivities, keffr and keffz,

respectively, and the wall heat transfer coefficient, hw. Namely, these parameters are

estimated using the boundary layer approximation that makes use of a two-dimensional

heat transport model coupled to the Navier-Stokes-Darcy-Forchheimer equation and the

existing void fraction profile [45].

Table 3 presents the main variables used for all simulations displayed in this study for

the industrial-scale reactor modeling. Reactor and catalyst dimensions are the same as those

used to obtain observations in absence of reaction from an industrial scale wall-cooled

packed bed reactor [68]. These pure heat transport experiments were used in a previous

research [45] to estimate the heat transport parameters (keffr, keffz, and hw.), which are crucial

for modeling the studied reactor with low dt/dp. Besides, the calculated active catalyst

density (ρb) is ca. 75 kg m-3, this value corresponds to the active MoVTeNbO formulation

coated over the external surface of TiO2 pellets packed in the reactor. The MoVTeNbO

active layer is considered to be 0.15 mm since mass and heat transport limitations are,

properly, minimized as observed from kinetic experiments. The feedstock is a mixture of

ethane, oxygen and an inert with composition defined from kinetic experiments. Thus, the

feed inlet O2/C2H6 ratios are strategically selected to be out of the explosive region.

Table 3 is shown here.

The wall-cooled packed bed reactor model is based on averaged general conservation

relations for mass and energy but considering effective transport parameters. The two-

dimensional pseudo-heterogeneous model used is given as follows:

15
Gas phase:

∂Cn ∂Cn  ∂ 2Cn 1 ∂Cn  ∂2 Cn


ε + uo = εDeffr  2 +  + εDeffz + (1 − ε)k g a s (Cns − Cn ) (24)
∂t ∂z  ∂r r ∂r  ∂z2

∂T ∂T  ∂ 2T 1 ∂T  ∂ 2T
ερf Cpf + uoρf Cpf = k effr  2 +  + keffz 2 + (1 − ε)hg as (Ts − T) (25)
∂t ∂z  ∂r r ∂r  ∂z

Solid phase:

∂Cns 5
(1 − ε) = (1 − ε ) k g a s (Cn − Cns ) + ρb ∑ ν ni ri (26)
∂t i =1

∂Ts 5
ρbCps = (1 − ε ) h g a s (T − Ts ) + ρb ∑ (−∆Hi )ri (27)
∂t i =1

The corresponding initial and boundary conditions are:

t = 0; Cn = Cn,ss and Cns = Cns,ss (28)

T = Tss and Ts = Ts,ss


(29)

∂Cn
z = 0; uoCno = uoCn − εDeffz (30)
∂z

∂T
uoρf Cpf To = uoρf Cpf T − keffz (31)
∂z

∂Cn ∂T
z = L; =0 and =0 (32)
∂z ∂z

∂Cn ∂T
r = 0; =0 and =0 (33)
∂r ∂r

16
∂Cn ∂T
r = Rt ; =0 and keffr = h w (T − Tb ) (34)
∂r ∂r

where Cn is the molar concentration of component n in the gas phase, Deffr is the radial

mass dispersion coefficient, Deffz is the axial mass dispersion coefficient, ρb is the fixed-bed

density (active catalyst density), ρf is the fluid density, Cpf is the specific heat capacity of

the fluid, Cps is the specific heat capacity of the solid, keffr is the radial effective thermal

conductivity, keffz is the axial effective thermal conductivity and hw is the wall heat transfer

coefficient.

The performance of the reactor is evaluated by the conversion of reactants as well as

by the yield of products. The conversion is defined as:

number of moles of n reacted


Xn = n = C2 H6 or O2 (35)
number of moles of n fed

while the yield of the n-th product is defined as:

number of moles of product n formed


Yn = n = C2 H 4 , CO, CO2 , and H 2O (36)
number of moles of ethane fed

The resulting model is given in terms of a set of parabolic partial differential

equations, which is solved numerically by the method of orthogonal collocation using 5 and

50 interior collocation points at the radial and axial coordinates, respectively, employing

shifted Legendre polynomials [70]. The resulting set of ordinary differential equations is,

then, solved by a Runge-Kutta method [71].

17
4 Results and discussion

The results are divided in two sections. Section 4.1 discusses the kinetics of ODH-Et on

MoVTeNbO catalytic system. The model reliability is firstly discussed, in terms of the

statistical significance of both the regression and the parameters; thus, the parameter values

are analyzed to evaluate whether they are physically meaningful or not. An analysis on the

adsorption phenomena of both water and oxygen on the active sites from MoVTeNbO

catalytic formulation is carried out in order to support the main kinetic assumptions. On the

other hand, Section 4.2 presents the industrial-scale catalytic reactor simulations, which are

aimed at elucidating the performance of MoVTeNbO catalyst during the ODH-Et in the

wall-cooled packed bed reactor. Firstly, the catalytic reactor is modeled either considering

those HTP that are estimated assuming plug flow, as treated nowadays in literature

simulations [68,72-74], or considering HTP that accounts for the effect of hydrodynamics

on heat transport phenomena [33,34,45]. The comparison of these simulations allows

elucidate the need for considering the role of velocity field on heat transport via the

corresponding HTP when a packed bed reactor presenting a low dt/dp is modeled. Secondly,

a parametric sensibility analysis is presented with the aim to unfold the influence of coolant

temperature, ethane and oxygen inlet concentrations on the catalytic performance of

MoVTeNbO catalyst. Finally, an analysis of the influence of the adsorption of both water

and oxygen on industrial reactor performance is presented.

4.1 Kinetics

MoVTeNbO material is an attractive catalytic formulation since it leads to ethane

conversion and selectivity to ethylene relevant for the industry [8]. A previous publication

18
[57] was dedicated to study the effect of feedstock composition, temperature and space-

time on a set of catalytic responses over the MoVTeNbO catalyst formulation. Fig. 1

summarizes main ODH-Et observations at different ethane conversions and selectivity to

ethylene, CO and CO2. Specifically, the selectivity to the reaction products decreases in the

following order: ethylene >> CO > CO2, ranging from 76 to 96 % for ethylene, 2.5 to 16.5

% for CO and 1.5 to 7.5 % for CO2. The region of high ethane conversion demands

operating at large values of both: temperature (480 ºC) and space-time (70 gcat h molethane-
1
), a scenario at which the selectivity to ethylene decreases and, hence, that of COx is

favored. Besides, Fig. 1 also presents how the proposed kinetic model (eqs. (6) - (11)) fits

adequately the conversion of ethane and the main reaction products in the ODH-Et.

Figure 1 is shown here.

The model adequacy is also evaluated by means of both statistical and

phenomenological tests. Fig. 2 shows the parity plot for the molar flow rates of all observed

components (C2H6, C2H4, CO, CO2, O2 and H2O). Model regression presents statistical

significance, yielding adequate agreement with the experimental data over a wide range of

experimental conditions, vide Table 1. Namely, deviations from observations are random

without presenting any undesired statistical trend. In this result, the random error is

constant, without the presence of a systematic association between experimental points at

any of the different operating conditions studied.

Figure 2 is shown here.

19
Table 4 presents the F-value of the regression, and the estimated parameters along

with the 95% confidence interval and t-value. The calculated F-value amounts to 7642, a

value three orders of magnitude larger than the tabulated one (2.79), corroborating the

satisfactory global significance of the model regression. Besides, all parameters are

statistically significant since their confidence interval is significantly narrow and their t-

value (ca. 240-435) is higher than the t-value tabulated (1.97). All parameters are not

statistically correlated since their binary correlations coefficients (ρij) are always lower than

0.5.

Table 4 is shown here.

As commented earlier the LHHW model has been the most suitable one nowadays for

describing the ODH-Et on the MoVTeNbO catalytic formulation [57], since other kinetic

models present a lack of statistical significance. In this regard, a statistical analysis is

carried out to discriminate between this previously reported LHHW model and the kinetic

model developed in this study. Notwithstanding the proposed kinetic model presents a

slightly better F-value (7642) than the LHHW model (7567), both models are evaluated

through the Bayesian information criterion (BIC) [79], in order to discard that model

resulting in overfitting because of the addition of unnecessary phenomena/parameters.

From this end, the developed kinetics is the most suitable one to describe ODH-Et

observations since it leads to a lower value of BIC (110.7) than the LHHW model (159.6),

obtaining an absolute difference (∆BIC) of ca. 49 [79]. Experimental data reported

elsewhere [57] elucidated how the inlet partial pressure of ethane gives rise to a positive

20
effect on ethane conversion due to its over oxidation to carbon oxides rather than its

adsorption on catalytic active sites. In this respect, the developed kinetics accounts for the

effect of inlet partial pressure of ethane on reaction rates; a situation that is not possible

when the LHHW model since it was insensitive to the inlet partial pressure of ethane [57].

Therefore, statistical and experimental evidences do indicate that the kinetic model

developed in this study is the most suitable to be evaluated during industrial-scale reactor

simulations.

The estimated values for the adsorption equilibrium constants for oxygen and water,

vide Table 4, indicate that standard adsorption enthalpies are negative with water being the

component with the smallest standard adsorption enthalpy ( ∆ H on = - 128.2 kJ mol-1). This

suggests that water binds more strongly to MoVTeNbO catalyst formulation than oxygen

( ∆ H on = - 45.6 kJ mol-1). Furthermore, Boudart´s criteria are totally satisfied since the

values of standard adsorption entropies are between 41.8 J (mol K)-1 and the corresponding

gas phase molecular standard entropy. The magnitude of the estimated standard adsorption

enthalpies is similar to those values reported in literature [57,75-77]. For instance, the

standard adsorption enthalpy of oxygen is reported from -19 to -98 kJ mol-1 [75] and the

adsorption heat of water, measured on various metal oxides surfaces, ranges between -36

and -113 kJ mol-1 [76,77]. According to the estimated activation energies, the rate of

formation of ethylene from ethane, vide Eq. (1), is the reaction that presents the lowest

activation energy i.e., 90.5 kJ mol-1. In contrast, the formation of COX from both ethane and

ethylene, vide eqs. (2)-(5), require activation energies ranging from 131.8 kJ mol-1 to 165.0

kJ mol-1. Specifically, ethylene is the principal contributor to COX because total oxidations

21
from ethane (ca. 150 kJ mol-1) require even more energy than the total oxidations from

ethylene (ca. 130 kJ mol-1). Our estimated values of activation energies, ranging from 90.5

to 165.0 kJ mol-1, are similar to the values reported in the literature for different catalytic

systems for the ODH-Et (from 50 to 150 kJ mol-1) [53,57,78]. The reaction orders

associated to the partial pressure of oxygen are found to be below 1.0, indicating that

reaction rates are weakly affected by changes in oxygen partial pressure.

4.1.1 Fractional surface coverage

Adsorption phenomenon on MoVTeNbO formulation seems a key aspect to be analyzed

during a proper kinetic study since reactant and product adsorptions affect significantly

reaction rates and, hence, catalyst activity and selectivity. Fig. 3 displays the predicted

surface coverage of adsorbed oxygen and adsorbed water as a function of temperature and

space-time, which are calculated from Eqs. (12) and (13), respectively. The MoVTeNbO

active sites are mostly covered with adsorbed water (ca. 0.95) rather than with adsorbed

oxygen (ca. 0.0). Thus, the present kinetics accounts properly for the negative effect of

adsorbed water on reaction rates since the high levels of water concentration and its affinity

on active sites, based on the highest water adsorption enthalpy, do indeed inhibit oxidative

dehydrogenation and total oxidations. Besides, the low fraction coverage of oxygen on

catalyst active sites and the low reaction orders associated to the partial pressure of oxygen,

indicate that the kinetics accounts satisfactorily for the weak effect of the inlet partial

pressure of oxygen on ethane conversion and selectivity to ethylene [57].

Figure 3 is shown here.

22
The presence of oxygen is, as observed, crucial for the oxidative dehydrogenation of

ethane and total oxidations to occur. Although catalytic studies [80,81] over V materials for

other reactions suggest that bulk lattice oxygen participates in the reaction and it is replaced

by catalyst reoxidation with oxygen from the gas phase, for the current catalyst the

superficial lattice oxygen seems to be the main species participating on several reactions

[14,57,81-84], which explains the high selectivity to ethylene. The activity of V catalysts

mainly depends on the oxygen’s lability, nevertheless this lability is a function of catalyst

chemical composition and crystalline structure [14,57,82,83]. In this respect, MoVTeNbO

formulation presents a relatively large capacity for oxygen recombination in its active M1

phase [14,57,82,83], i.e. the oxygen species are immediately used for oxidation reactions

from the first lattice layers and are overtaken from the gas phase. This might be the reason

why the proposed mechanism, vide Table 2, adequately represents the observed kinetic

phenomenon for MoVTeNbO rather the well-known Mars and van-Krevelen redox

mechanism [85]. Kubo et al., [84], during the evaluation of propane ammoxidation with

lattice oxygen from Mo-V-O-based complex metal oxide catalysts, observed that the

migration rate of the oxygen from the bulk to superficial lattice is too slow in comparison

with the consumption rate of oxygen from the surface lattice. In this respect, a relatively

small amount of oxygen is demanded to reoxidize the metallic sites that are reduced after

the propane oxidation reactions, thus, explaining the high selectivity to ethylene of this type

of mixed oxide catalysts. Namely, a low amount of oxygen on surface lattice along with a

low reaction temperature give rise to ethylene formation out of ethane rather than total

oxidations.

23
4.2 Industrial-scale reactor predictions

The adequate consideration of heat transport phenomena in the model of a wall-cooled

packed bed catalytic reactor with low dt/dp is, undoubtedly, one of the most essential

aspects to be accounted for during the conceptual design of this type of reaction systems.

There are several approximations that have been used to model heat transport in a wall-

cooled catalytic reactor with low dt/dp. Some studies [33,34,45] suggest that the inclusion

of velocity profiles improves heat transport description; however, others studies [68,72-74]

have neglected the influence of the hydrodynamics in the modeling of this type of reactors.

In this regard, a practical but reliable approach, developed by our research group, is herein

applied to account for the role of hydrodynamics on heat transport during the modeling of

the ODH-Et in a wall-cooled catalytic reactor packed with MoVTeNbO catalyst externally

deposited on a nonporous support of TiO2. Particularly, the boundary layer approximation

is used to estimate keff and hw from heat transfer observations in absence of reaction, thus,

these parameters containing the hydrodynamic information are used to model the industrial

scale reactor by means of the classical model, vide eqs. (24)-( 34). This approach uses a

less expensive numerical scheme since it only contains heat and mass balances instead of

momentum, heat and mass balances which computationally speaking is very time-

consuming.

In order to be judicious on the modelling approach applied to elucidate the role of

MoVTeNbO externally deposited on TiO2 for the ODH-Et in a wall-cooled catalytic reactor

with low dt/dp packed, we present the simulations of an industrial-scale reactor for the

catalytic o-xylene partial oxidation using the aforementioned approach. The o-xylene

partial oxidation was selected since observations at industrial-scale with and without

24
reaction are available from studies carried out by the research group [33, 34, 68]. Moreover,

this reaction system presents the same geometric (L and dt/dp) and operational conditions

(Rep) that are used for the ODH-Et reaction in this work, vide Table 3. From this end, Fig. 4

display temperature profiles for two oxidation reactions, let´s say, the o-xylene partial

oxidation used herein as a reference (Fig. 4a), and the ODH-Et (Fig. 4b). Both reaction

cases compare temperature profiles along the reactor using distinct effective HTP estimated

from the same set of experiments in absence of reaction: i) parameters estimated via the

classical approach (CA) assuming an uniform velocity profile with constant void fraction

[68,72-74] and ii) parameter estimated via the boundary layer approximation (BLA) in

which the role of void fraction profile and velocity field are assessed [33,34,45]. Table 3

presents the numerical values of hw and keff estimated via the BLA and CA from the same

experimental dataset. These HTP are larger for BLA, reflecting differences in the estimated

rates of radial heat transport through the core of the packed bed and close to the wall. To

this end, the predictions of measured temperature profiles in Fig. 4a for the o-xylene partial

oxidation case elucidate the suitability of the HTP values from the BLA, giving better

predictions than when HTP values using the CA approach is employed in the reactor

model, vide, i.e. eqs. (24)-( 34). Therefore, hw and keff are effective parameters that through

BLA lump differently the hydrodynamic and heat-transport phenomena in the packed bed

leading to better predictions. Fig. 4b present temperature profiles for the ODH-Et. As

expected, when the reactor model makes use of HTP estimated considering plug flow, there

is the presence of a higher hot spot since the role of hydrodynamics on heat transfer

mechanisms is neglected. Owing to the above, the following sections present the

simulations of the industrial-scale reactor for ODH-Et considering those HTP estimated

from BLA.
25
Figure 4 is shown here.

4.2.1 Parametric sensitivity study

The behavior of a reactor is a function of the initial process conditions, reactor disturbances

and the trajectories of the manipulated reactor variables. In order to analyze an open-loop

parametric sensitivity study for the performance of MoVTeNbO system during the ODH-Et

in an industrial scale wall-cooled packed bed reactor, this section is aimed at simulating the

reactor when perturbations of cooling bath temperature and inlet ethane concentration are

considered. The first case, illustrated in Fig. 5, shows the predicted dynamic response when

reactor starts up at Tb of 400ºC with an inlet molar ratio C2H6/O2/N2 = 9/7/84, once the

steady state is reached, it is followed by perturbations of the cooling bath temperature,

namely 440ºC and 480ºC. The second case, illustrated in Fig. 6, displays the predicted

dynamic response when reactor starts up at Tb of 400ºC with an inlet molar ratio of

C2H6/O2/N2 = 1/7/92, once the steady state is reached, it is followed by perturbations of

the inlet molar ratio of ethane, specifically C2H6/O2/N2 = 9/7/84, C2H6/O2/N2 = 18/7/75

and C2H6/O2/N2 = 40/7/53.

Fig. 5 displays transient and steady state temperature, conversion (ethane and oxygen) and

yield (ethylene, COX and water) profiles along the reactor length at different coolant

temperatures maintaining a fixed inlet molar ratio C2H6/O2/N2=9/7/84. So far, an increment

in coolant temperature causes an increase in the magnitude of the hot spot, vide Fig. 5a.

Specifically a Tb equal to 400 ºC and 440 ºC leads to mild hot spots of ca. 1 ºC and 7 ºC

respectively whereas a Tb equal to 480 ºC provokes a pronounced hot spot with a


26
temperature rise of ca. 56 ºC. In the latter, the rate of heat removal is minor in comparison

with the rate of heat generated. Besides Figs. 5b-d displays the effect of coolant

temperature on the overall consumption rates of the reactants. The increment of Tb favors

both ethane and oxygen conversions. Analyzing the product distribution, an increase in Tb

brings out a positive effect on the yield of ethylene out of ethane. However, a higher Tb

leads to higher carbon oxides yields, which are the responsible of both the formation of hot

spots and the decrease in the selectivity to ethylene. In this regard, a lower Tb is

recommended for a satisfactory ethane conversion along with ethylene selectivity avoiding

catalyst damage due to high temperatures [23]. The behavior of the industrial scale reactor

presented in Fig. 5 shows similar tendencies as those of the typical industrial-scale packed

bed reactors used to perform highly exothermic reactions [24,30,33,34,67,68].

Figure 5 is shown here.

In the design of this type of complex reactors it is necessary to operate without the

presence of drastic hot spots in order to decrease the irreversible deactivation of the

catalysts by a sintering phenomenon [86]. Particularly, the stability of the MoVTeNbO

catalyst is restricted to operate below 500 °C, due to energy saving but, mainly, because

operations above 500 °C leads to the irreversible loss of catalytic activity as a consequence

of the removal of tellurium from the MoVTeNbO catalytic formulation [23], i.e.

specifically, the lost of Te from MoVTeNbO material occurs preferentially from the end

section of the M1 crystal phase, across the [001] plane. The removal of Te species modifies

the crystal phase composition of the M1 phase causing the partial destruction of the

27
MoVTeNbO with the simultaneous formation of the MoO2 phase and, in consequence, the

activity and selectivity to ethylene are affected as elucidated elsewhere [23].

Fig. 6 compares transient and steady state temperature, conversion (ethane and

oxygen) and yield (ethylene, COX and water) profiles along the reactor length at different

ethane inlet concentrations at Tb equals to 440 ºC. The position of the hot spot is located

toward the first meter of reactor length, a zone with higher concentration of ethane. An

increment in the inlet ethane concentration causes a slight increment of the magnitude of

the hot spot, vide Fig. 6a. Particularly, 1% or 9 % of ethane in the feedstock leads to mild

hot spots of ca. 1 ºC or 7 ºC, respectively, whereas 18% or 40% of ethane causes hot spots

with a temperature rise of ca. 12 ºC or 34 ºC, respectively. Figs. 6b-e shows the effect of

inlet ethane concentration on ethane conversion and product yields. An increase in ethane

inlet concentration causes a rapid oxydehydrogenation (r1), but mainly of the total

oxidations (r2-r5) reactions, thus, provoking large amounts of heat generation. Therefore,

the increase in ethane inlet concentration favors reactant conversions and product yields

rather than ethylene selectivity (vide Figs. 6b-e).

Figure 6 is shown here.

Finally, a brief analysis of the influence of the adsorption of both water and oxygen

on catalytic reactor performance is presented. As can be observed in Fig. 7, the catalyst

surface is mainly occupied by water and vacancy sites, since the surface coverage of

oxygen is found to be negligibly small along the reactor length, as elucidated from

laboratory experiments in Section 4.1.1. The reactions occurring on the MoVTeNbO

28
catalytic formulation are a function of the surface coverage profiles. The relatively low

fraction coverage of oxygen determines the product distribution along the reactor, revealing

the selective role of the MoVTeNbO catalyst to produce ethylene out of ethane. Aside, the

high fraction coverage of water leads to a negative effect on both type of reactions, those

selective to produce ethylene and those selective to obtain COx (vide Fig. 5). Therefore,

water inhibits the oxidative dehydrogenation of ethane and total oxidations through its

adsorption on MoVTeNbO sites. Thus, water seems to play an important role on this type

of catalytic reaction, as documented for other catalytic partial oxidations of different

hydrocarbons [87].

Figure 7 is shown here.

5 Conclusions

The implementation of a new catalytic material at industrial level requires the knowledge

from the conceptual design of the reactor technology. Modeling has been the most reliable

engineering tool to carry out the preliminary reactor design. In this regard, this work shows

through modeling the catalytic performance of MoVTeNbO catalyst for ODH-Et in a wall-

cooled industrial-scale catalytic packed bed reactor presenting a low dt/dp ratio. To have a

confidence on model simulations, a proper determination of both kinetic parameters and

heat transport parameters from independent experiments should be carefully carried out. In

terms of the kinetics, an adequate agreement between model and experimental values is

observed, wherein the obtained parameters are physically meaningful and are found to be

statistically significant. The proposed kinetics addresses several aspects observed

29
experimentally, namely it accounts for both the negative effect of adsorbed water on

reaction rates and the weak effect of the inlet partial pressure of oxygen on ethane

conversion and selectivity to ethylene. In this respect, even the presence of oxygen is

crucial for the oxidative dehydrogenation of ethane and total oxidations to occur, the

superficial lattice oxygen from the MoVTeNbO catalyst seems to be the main species

participating on reactions, explaining, thus, its high selectivity to ethylene.

In terms of the industrial-scale reactor, the simulations suggest that hydrodynamics is

an essential aspect to be considered during the modeling of this type of reaction systems

presenting a low dt/dp. The coolant temperature is the main variable affecting ethane

conversions and yield to ethylene. For instance, coolant temperatures larger than 440ºC

lead to hot spot magnitudes that could cause the damage to the MoVTeNbO structure. As

observed from laboratory experiments, water formation affects conversion and selectivity,

and catalyst is weakly oxidized along the reactor length, giving insights on the highly

selectivity of the MoVTeNbO catalytic formulations.

On the basis of the results obtained herein, the wall-cooled packed-bed reactor

represents a promising alternative for petrochemical industry to carry out the ODH-Et over

a MoVTeNbO catalyst, which is highly selective to ethylene. Besides, the model

developed in this work, in general, and the modeling strategy based on laboratory and pilot

plant experiments in absence of reaction, in particular, are potential tools to be used in

future studies for the conceptual design and scale-up of industrial reactor system for the

ODH-Et on MoVTeNbO material.

30
Acknowledgments

This study was supported by El Consejo Nacional de Ciencia y Tecnología (CONACyT)

under project No. 181104. Gamaliel Che-Galicia also thanks CONACyT for providing a

postgraduate fellowship. We also acknowledge to Eng. J.C. Castillo-Rodriguez, Eng. C.

Tzompantzi-Flores and Eng. O. Romero-Pérez for the technical support and discussion on

the developed kinetic model. Authors also thank valuable comments from the anonymous

reviewers whose careful revisions and critics decisively contributed in improving the

quality of this paper.

Nomenclature

Roman letters

as external surface to particle volume ratio, ms-1

A 'i natural logarithm of pre-exponential factor, mmol (gcat h)-1

Cn molar concentration of component n in the gas phase, kmol mf-3

Cn,ss molar concentration of component n in the gas phase at steady state, kmol mf-3

Cno feed molar concentration of component n, kmol mf-3

Cns molar concentration of component n on the solid phase, kmol ms-3

Cns,ss molar concentration of component n on the solid phase at steady state, kmol ms-3

Cpf heat capacity of the fluid, kJ (kgf K)-1

Cps heat capacity of the solid, kJ (kgs K)-1

dp particle diameter, ms

dt reactor diameter, mr

31
Deffr radial mass dispersion coefficient, mr2 h-1

Deffz axial mass dispersion coefficient, mr2 h-1

EA activation energy, kJ mol-1

Fn molar flow rate of the component n, mmol h-1

hg interfacial heat transport coefficient, kJ ms-2 (h K)-1

hw wall heat transfer coefficient, kJ mr-2 (h K)-1

ki reaction rate coefficient, mmol (gcat h)-1

keffr radial effective thermal conductivity, kJ (mr h K)-1

keffz axial effective thermal conductivity, kJ (mr h K)-1

kg interfacial mass transfer coefficient, mf3 ms-2 h-1

Kn adsorption equilibrium coefficient for component n, Pa-1

L reactor length, mr

m partial reaction order for oxygen

pn partial pressure of component n, Pa

r radial position, mr

ri specific reaction rate of reaction i, mmol (gcat h)-1

Rn net reaction rate of the component n, mmol (gcat h)-1

R universal gas constant, kJ (mol K)-1

RSS objective function (residual sum of squares)

S active sites for kinetic model mechanism

Sg standard entropy of gas phase molecule, J (mol K)-1

t time, h

T temperature, K

To feed temperature, K

Tb salt bath temperature, K

32
Ts solid temperature, K

Ts,ss steady state solid temperature, K

Tss steady state fluid temperature, K

uo superficial velocity, mf3 mr-2 h-1

wn objective function weight factor of each response

Wcat mass of catalyst, gcat

X n conversion of component n

Yn yield of component n

z axial position, mr

Greek letters

α vector of parameters accounted for in the objective function

β vector of parameters accounted for in the objective function

∆ H on standard enthalpy of adsorption for component n, kJ (mol K)-1

∆H i reaction enthalpy, kJ mol-1

∆Son Standard entropy of adsorption for component n used in Eq. (13), kJ (mol K)-1

∆Son Standard entropy of adsorption for component n, J (mol K)-1

γ C2 H6 ratio of oxygen and ethane partial pressure, mol O 2 mol −C12 H 6

γ C2 H 4 ratio of oxygen and ethane partial pressure, mol O 2 mol −C12 H 4

ε void fraction, mf3 mr-3

νi stoichiometric number

33
θ fractional surface coverage

ρb bed density, kgcat mr-3

ρf fluid density, kgf mf-3

σj Horiuti number

Subscripts

cat catalyst

eff effective

exp experiment

f fluid

n component n

o inlet, superficial

p particle

r reactor

s solid

tab tabulated

* vacant sites density

Superscripts

^ calculated

o inlet, standard

* reference

34
References

[1] A. Clements, M. Dunn, V. Firth, L. Hubbard, J. Lazonby, D. Waddington, The

essential chemical industry, Chemical Industry Education Centre, York, 2010.

[2] M.A. Bañares, Supported metal oxide and other catalysts for ethane conversion: a

review, Catal. Today 51 (1999) 319–348.

[3] S. Albonetti, F. Cavani, F. Trifirò, Key aspects of catalyst design for the selective

oxidation of paraffins, Catal. Rev. 38 (1996) 413–438.

[4] F. Cavani, F Trifirò, The oxidative dehydrogenation of ethane and propane as an

alternative way for the production of light olefins, Catal. Today 24 (1995) 307–313.

[5] C.A. Gärtner, A.C. van Veen, J.A. Lercher, Oxidative Dehydrogenation of Ethane:

Common Principles and Mechanistic Aspects, ChemCatChem 5 (2013) 3196–3217.

[6] K. Weissermel, H.-J. Arpe, Industrial Organic Chemistry, WILEY-VCH Verlag

GmbH & Co. KGaA, Weinheim, 2003.

[7] T. Blasco, J.M. López Nieto, Oxidative dehydrogenation of short chain alkanes on

supported vanadium oxide catalysts, Appl. Catal. A: Gen. 157 (1997) 117–142.

[8] F. Cavani, N. Ballarini, A. Cericola, Oxidative dehydrogenation of ethane and

propane: how far from commercial implementation?, Catal Today 127 (2007) 113–131.

[9] E.M. Thorsteinson, T.P. Wilson, F.G. Young, P.H. Kasai, The Oxidative

Dehydrogenation of Ethane over Catalysts Containing Mixed Oxides of Molybdenum and

Vanadium, J. Catal. 52 (1978) 116–132.

35
[10] E. Morales, J. H. Lunsford, Oxidative dehydrogenation of ethane over a lithium-

promoted magnesium oxide catalyst, J. Catal. 118 (1989) 255–265.

[11] J.M. López Nieto, P. Botella, M.I. Vázquez, A. Dejoz, The selective oxidative

dehydrogenation of ethane over hydrothermally synthesized MoVTeNb catalysts, Chem.

Commun. (2002) 1906–1907.

[12] M. Panizza, C. Resini, F. Raccoli, G. Busca, R. Catani, S. Rossini, Oxidation of

ethane over vanadia-alumina-based catalysts: co-feed and redox experiments, Chem. Eng.

J. 93 (2003) 181–189.

[13] P. Botella, E. García-González, A. Dejoz, J.M. López Nieto, M.I. Vázquez, J.

González-García, Selective oxidative dehydrogenation of ethane on MoVTeNbO mixed

metal oxide catalysts, J. Catal. 225 (2004) 428–438.

[14] P. Botella, E. García-González, J.M. López Nieto, J.M. González-Calbet,

MoVTeNbO multifunctional catalysts: Correlation between constituent crystalline phases

and catalytic performance, Solid State Sci. 7 (2005) 507–519.

[15] E. Heracleous, A. A. Lemonidou, Ni-Nb-O mixed oxides as highly active and

selective catalysts for ethene production via ethane oxidative dehydrogenation. Part I:

Characterization and catalytic, J. Catal. 237 (2006) 162–174.

[16] B. Solsona, M.I. Vázquez, F. Ivars, A. Dejoz, P. Concepción, J.M. López Nieto,

Selective oxidation of propane and ethane on diluted Mo-V-Nb-Te mixed-oxide catalysts,

J. Catal. 252 (2007) 271–280.

36
[17] X. Lin, C.A. Hoel, W.M.H. Sachtler, K.R. Poeppelmeier, E. Weitz, Oxidative

dehydrogenation (ODH) of ethane with O2 as oxidant on selected transition metal-loaded

zeolites, J. Catal. 265 (2009) 54–62.

[18] J.S. Valente, R. Quintana-Solórzano, H. Armendáriz-Herrera, G. Barragán-

Rodríguez, J.M. López-Nieto, Kinetic Study of Oxidative Dehydrogenation of Ethane over

MoVTeNb Mixed-Oxide Catalyst, Ind. Eng. Chem. Res. 53 (2014) 1775–1786.

[19] J. Santander, E. López, A. Diez, M. Dennehy, M. Pedernera, G. Tonetto, Ni-Nb

mixed oxides: One-pot synthesis and catalytic activity for oxidative dehydrogenation of

ethane, Chem. Eng. J. 255 (2014) 185–194.

[20] J.M. López Nieto, P. Botella, B. Solsona, J.M. Oliver, The selective oxidation of

propane on Mo-V-Te-Nb-O catalysts: the influence of Te-precursor, Catal. Today 81

(2003) 87–94.

[21] P. Concepción, P. Botella, J.M. López Nieto, Catalytic and FT-IR study on the

reaction pathway for oxidation of propane and propylene on V- or Mo-V based catalysts,

Appl. Catal. A: Gen. 278 (2004) 45–56.

[22] B. Solsona, F. Ivars, P. Concepción, J.M. López Nieto, Selective oxidation of n-

butane over MoV-containing oxidic bronze catalysts, J. Catal. 250 (2007) 128–138.

[23] J.S. Valente, H. Armendáriz-Herrera, R. Quintana-Solórzano, P. del Ángel, N. Nava,

A. Massó, J.M. López Nieto, Chemical, structural, and morphological changes of a

MoVTeNb catalyst during oxidative dehydrogenation of ethane, ACS Catal. 4 (2014)

1292–1301.

37
[24] G.F. Froment, K.B. Bischoff, Chemical reactor analysis and design, John Wiley &

Sons, New York, 1979.

[25] P. Arpentinier, F. Cavani, F. Trifiro, The Technology of Catalytic Oxidations,

Technip, Paris, 2001.

[26] J. Lerou, G.F. Froment, Velocity, Temperature and Conversion Profiles in Fixed Bed

Catalytic Reactor, Chem. Eng. Sci. 32 (1977) 853–861.

[27] T. Daszkowski, G. Eigenberger, A Reevaluation of Fluid Flow Heat Transfer and

Chemical Reaction in Catalyst Filled Tubes, Chem. Eng. Sci. 47 (1992) 2245–2250.

[28] V. Blasco, C. Royo, A. Monzbn, J. Santamaria, Catalyst Sintering in Fixed-Bed

Reactors: Deactivation Rate and Thermal History, AIChE J. 38 (1992) 237–243.

[29] Y.S. Cheng, F. López-Isunza, T. Mongkhonsi, L.S. Kershenbaum, Estimation of

Catalyst Activity Profiles in Fixed-Bed Reactors with Decaying Catalysts, Appl. Catal. A

106 (1993) 193–199.

[30] J.N. Papageorgiou, G.F. Froment, Simulation Models Accounting for Radial Voidage

Profiles in Fixed-Bed Reactors, Chem. Eng. Sci. 50 (1995) 3043–3056.

[31] O. Bey, G. Eigenberger, Gas Flow and Heat Transfer through Catalyst Filled Tubes,

Int. J. Therm. Sci. 40 (2001) 152–164.

[32] D. Wen, Y. Ding, Heat Transfer of Gas Flow through a Packed Bed, Chem. Eng. Sci.

61 (2006) 3532–3542.

[33] C.O. Castillo-Araiza, F. López-Isunza, Modeling the Partial Oxidation of o-Xylene in

an Industrial Packed-Bed Catalytic Reactor: The Role of Hydrodynamics and Catalyst

Activity in the Heat Transport, Ind. Eng. Chem. Res. 49 (2010) 6845–6853.

38
[34] C.O. Castillo-Araiza, F. López-Isunza, The role of catalyst activity on the steady state

and transient behavior of an industrial-scale fixed bed catalytic reactor for the partial

oxidation of o-xylene on V2O5/TiO2, Chem. Eng. J. 176-177 (2011) 26–32.

[35] M. Fattahi, M. Kazemeini, F. Khorasheh, A. Darvishi, A.M. Rashidi, Fixed-bed

multi-tubular reactors for oxidative dehydrogenation in ethylene process, Chem. Eng.

Technol. 36 (2013) 1692–1700.

[36] Z. Skoufa, E. Heracleous, A.A. Lemonidou, Investigation of engineering aspects in

ethane ODH over highly selective Ni0.85Nb0.15Ox catalyst, Chem. Eng. Sci., 84 (2012) 48–

56.

[37] M.L. Rodriguez, D.E. Ardissone, E. Heracleous, A.A. Lemonidou, E. López, M.N.

Pedernera, D.O. Borio, Oxidative dehydrogenation of ethane to ethylene in a membrane

reactor: A theoretical study, Catal. Today 157 (2010) 303–309.

[38] M.L. Rodriguez, D.E. Ardissone, A.A. Lemonidou, E. Heracleous, E. López, M.N.

Pedernera, D.O. Borio, Simulation of a membrane reactor for the catalytic

oxidehydrogenation of ethane, Ind. Eng. Chem. Res., 48 (2009) 1090–1095.

[39] C. Hamel, Á. Tóta, F. Klose, E. Tsotsas, A. Seidel-Morgenstern, Analysis of single

and multi-stage membrane reactors for the oxidation of short-chain alkanes-Simulation

study and pilot scale experiments. Chem. Eng. Res. Des., 86 (2008) 753–764.

[40] E. López, E. Heracleous, A.A. Lemonidou, D.O. Borio, Study of a multitubular fixed-

bed reactor for ethylene production via ethane oxidative dehydrogenation, Chem. Eng. J.,

145 (2008) 308–315.

39
[41] F.A. Al-Sherehy, A.M. Adris, M.A. Soliman, R. Hughes, Avoidance of flammability

and temperature runaway during oxidative dehydrogenation using a distributed feed, Chem.

Eng. Sci. 53 (1998) 3965–3976.

[42] R.B. Benenati, C.F. Brosilow, Void Fraction Distribution in Beds of Spheres, J. Am.

Inst. Chem. Eng. 8 (1962) 359–361.

[43] E. Tsotsas, E.U. Schlunder, Some Remarks on Channelling and on Radial Dispersion

in Packed Beds. Chem. Eng. Sci. 43 (1988), 1200–1203.

[44] M. Winterberg, E. Tsotsas, A. Krischke, D. Vortmeyer, A Simple and Coherent Set of

Coeficients for Modelling of Heat and Mass Transport with and without Chemical Reaction

in Tubes Filled with Spheres, Chem. Eng. Sci. 55 (979) 967–979.

[45] C.O. Castillo-Araiza, H. Jiménez-Islas, F. López-Isunza, Heat-transfer studies in

packed-bed catalytic reactors of low tube/particle diameter ratio, Ind. Eng. Chem. Res. 46

(2007) 7426–7435.

[46] C.O. Castillo-Araiza, F. Lopez-Isunza, Hydrodynamic models for packed beds with

low tube-to-particle diameter ratio, Int. J. Chem. Reactor Eng. 6 (2008) A1.

[47] S.T. Oyama, A.A. Middlebrook, G.A. Somorjai, Kinetics of Ethane Oxidation on

Vanadium Oxide, J. Phys. Chem. 94 (1990) 5029–5033.

[48] J. Le Bars, J.C. Vedrine, A. Auroux, Role of surface acidity on vanadia/silica

catalysts used in the oxidative dehydrogenation of ethane, Appl. Catal. A: Gen. 88 (1992)

179–195.

[49] C-Y. Kao, K-T. Huang, B-Z. Wan, Ethane Oxydehydrogenation over Supported

Vanadium Oxides, Ind. Eng. Chem. Res. 33 (1994) 2066–2072.

40
[50] M.D. Argyle, K. Chen, At.T Bell, E. Iglesia, Ethane Oxidative Dehydrogenation

Pathways on Vanadium Oxide Catalysts, J. Phys. Chem. B 106 (2002) 5421–5427.

[51] D. Linke, D. Wolf, M. Baerns, S. Zeyß, U. Dingerdissen, Catalytic Partial Oxidation

of Ethane to Acetic Acid over Mo1V0.25Nb0.12Pd0.0005Ox: II. Kinetic Modelling, J. Catal. 205

(2002) 32–43.

[52] F. Klose, M. Joshi, C. Hamel, A. Seidel-Morgenstern, Selective oxidation of ethane

over a VOx/γ-Al2O3 catalyst - investigation of the reaction network, Appl. Catal. A: Gen.

260 (2004) 101–110.

[53] R. Grabowski, J. Słoczyński, Kinetics of oxidative dehydrogenation of propane and

ethane on VOx/SiO2 pure and with potassium additive, Chem. Eng. Process. 44 (2005)

1082–1093.

[54] E. Heracleous, A. A. Lemonidou, Ni-Nb-O mixed oxides as highly active and

selective catalysts for ethene production via ethane oxidative dehydrogenation. Part II:

Mechanistic aspects and kinetic modeling, J. Catal. 237 (2006) 175–189.

[55] F. Rahman, K.F. Loughlin, M.A. Al-Saleh, M.R. Saeed, N.M. Tukur, M.M. Hossain,

K. Karim, A. Mamedov, Kinetics and mechanism of partial oxidation of ethane to ethylene

and acetic acid over MoV type catalysts, Appl. Catal. A: Gen. 375 (2010) 17–25.

[56] S.A. Al-Ghamdi, M.M. Hossain, H.I. de Lasa, Kinetic Modeling of Ethane Oxidative

Dehydrogenation over VOx/Al2O3 Catalyst in a Fluidized-Bed Riser Simulator, Ind. Eng.

Chem. Res. 52 (2013) 5235–5244.

41
[57] G. Che-Galicia, R. Quintana-Solórzano, J.S. Valente, R.S. Ruiz-Martínez, C.O.

Castillo-Araiza, Kinetic modeling of the oxidative dehydrogenation of ethane to ethylene

over a MoVTeNbO catalytic system, Chem. Eng. J. 252 (2014) 75–88.

[58] R. Quintana-Solórzano, G. Barragán-Rodríguez, H. Armendáriz-Herrera, J.M. López-

Nieto, J.S. Valente, Understanding the kinetic behavior of a Mo–V–Te–Nb mixed oxide in

the oxydehydrogenation of ethane, Fuel 138 (2014) 15–26.

[59] P.N. Brown, G.D. Byrne, A.C. Hindmarsh, VODE: a variable-coefficient ODE

solver, SIAM J. Sci. Stat. Comput. 10 (1989) 1038–1051.

[60] H.H. Rosenbrock, An automatic method for finding the greatest or least value of a

function, Comput. J. 3 (1960) 175–184.

[61] P.T. Boggs, J.R. Donaldson, R.H. Byrd, R.B. Schnabel, Algorithm 676 ODRPACK:

software for weighted orthogonal distance regression, ACM T. Math. Software 15 (1989)

348–364.

[62] D. .W. Marquardt, An algorithm for least-squares estimation of nonlinear parameters,

J. Soc. Ind. Appl. Math. 11 (1963) 431–441.

[63] E. Santacesaria, Kinetics and transport phenomena, Catal. Today 34 (1997) 393-400.

[64] C.H. Bartholomew, Mechanisms of catalyst deactivation, Appl. Catal. A: Gen. 212

(2001) 17–60.

[65] M. Boudart, D.E. Mears, M.A. Vannice, Kinetics of heterogeneous catalytic

reactions, Ind. Chim. Belg. 32 (1967) 281–284.

[66] G.F. Froment, Fixed Bed Catalytic Reactors. Technological and Fundamental Design

Aspects. Chemie Ingenieur Technik 46 (1974) 374–386.

42
[67] A.I. Anastasov, An investigation of the kinetic parameters of the o-xylene oxidation

process carried out in a fixed bed of high-productive vanadia-titania catalyst, Chem. Eng.

Sci. 58 (2003) 89–98.

[68] F. López-Isunza, Steady state and dynamic behavior of an industrial fixed bed

catalytic reactor, Ph.D. Thesis. University of London (Imperial College), United Kingdom,

1983.

[69] C.N. Satterfield, Mass Transfer in Heterogeneous Catalysis, M.I.T. Press, Cambridge,

1970.

[70] B. Finlayson, Nonlinear Analysis in Chemical Engineering, McGraw-Hill, New

York, 1980.

[71] L. Lapidus, J. Seinfeld, Numerical Solution of Ordinary Differential Equations,

Academic Press, New York, 1971.

[72] C. H. Li, B.A. Finlayson, Heat transfer in packed beds-a reevaluation, Chem. Eng.

Sci. 32 (1977) 1055–1066.

[73] B.D. Kulkarni, L.K. Doraiswamy, Estimation of effective transport properties in

packed bed reactors, Catal. Rev. 22 (1980) 431–483.

[74] D. Wen, Y. Ding, Heat transfer of gas flow through a packed bed, Chem. Eng. Sci. 61

(2006) 3532–3542.

[75] G. Saracco, F. Geobaldo, G. Baldi, Methane combustion on Mg-doped LaMnO3

perovskite catalysts, Appl. Catal. B: Environ. 20 (1999) 277–288.

43
[76] M. Egashira, S. Kawasumi, S. Kagawa, T. Seiyama, Temperature programmed

desorption study of water adsorbed on metal oxides. I. Anatase and rutile, Bull. Chem. Soc.

Jpn. 51 (1978) 3144–3149.

[77] M.P. Heynderickx, J.W. Thybaut, H. Poelman, D. Poelman, G.B. Marin, Kinetic

modeling of the total oxidation of propane over CuO-CeO2/c-Al2O3, Appl. Catal. B:

Environ. 95 (2010) 26–38.

[78] D. Wolf, N. Dropka, Q. Smejkal, O. Buyevskaya, Oxidative dehydrogenation of

propane for propylene production–comparison of catalytic processes, Chem. Eng. Sci. 56

(2001) 713–719.

[79] G. Schwarz, Estimating the Dimension of a Model, Ann. Stat. 6 (1978) 461-464.

[80] J.C. Vedrine, J.M.M. Millet, J-C. Volta, Molecular description of active sites in

oxidation reactions: Acid-base and redox properties, and role of water, Catal. Today 32

(1996) 115–123.

[81] K. Chen, A. Khodakov, J. Yang, A.T. Bell, E. Iglesia, Isotopic Tracer and Kinetic

Studies of Oxidative Dehydrogenation Pathways on Vanadium Oxide Catalysts, J. Catal.

186 (1999) 325–333.

[82] M. Aouine, J.L. Dubois, J.M.M. Millet, Crystal chemistry and phase composition of

the MoVTeNbO catalysts for the ammoxidation of propane, Chem. Commun. 13 (2001)

1180–1181.

[83] M. Hävecker, S. Wrabetz, J. Kröhnert, L-I. Csepei, R. Naumann d’Alnoncourt, Y.V.

Kolen’ko, F. Girgsdies, R. Schlögl, A. Trunschke, Surface chemistry of phase-pure M1

44
MoVTeNb oxide during operation in selective oxidation of propane to acrylic acid, J. Catal.

285 (2012) 48–60.

[84] J. Kubo, N. Watanabe, W. Ueda, Propane ammoxidation with lattice oxygen of Mo-

V-O-based complex metal oxide catalysts, Chem. Eng. Sci. 63 (2008) 1648–1653.

[85] P. Mars, D.W. van Krevelen, Oxidations carried out by means of vanadium oxide

catalysts, Chem. Eng. Sci. Spec. Suppl. 3 (1954) 41–59.

[86] S. S. E. H. Elnashaie, S. S. Elshishini, Modelling, Simulation and Optimization of

Industrial Fixed Bed Catalytic Reactors, Gordon and Breach Science Publishers, Yverdon,

1993.

[87] M.M. Bettahar, G. Costentin, L. Savary, J.C. Lavalley, On the partial oxidation of

propane and propylene on mixed metal oxide catalysts, App. Catal. A: Gen. 145 (1996) 1–

48.

45
Figure captions:

Figure 1. Selectivity to ethylene, CO2, and CO versus ethane conversion. Full lines are

calculated with the kinetic model (T = 400–480 ºC; space-time = 23–70 gcat h molethane-1;

nominal inlet molar ratio C2/O2/N2 = 9/7/84).

Figure 2. Parity plot comparing experimental with calculated reactor outlet molar flow

rates for the kinetic model. The full lines are the first bisector and the dashed lines represent

a deviation of 10 %.

Figure 3. (a) Fractional surface coverage versus temperature at inlet molar ratio

C2H6/O2/N2 = 9/7/84 and space-time of 70 gcat h molethane,o-1 ; and (b) Fractional coverage

versus space-time at inlet molar ratio C2H6/O2/N2 = 9/7/84 and T = 440 ºC.

Figure 4. Axial temperature profiles in the reactor using different HTP: (a) Comparison

between prediction and observation for the partial oxidation of o-xylene on a V2O5/TiO2

catalyst at: Tb = 376 °C, inlet molar ratio o-xylene/air = 1/99 and To = 190 ºC [33]; and (b)

Predictions for the oxidative dehydrogenation of ethane on MoVTeNbO/TiO2 catalyst at:

Tb = 440 °C, inlet molar ratio C2H6/O2/N2 = 9/7/84 and To = 200 ºC.

Figure 5. Reactor model predictions when Tb is increased 40 ºC in a ramp change with an

inlet molar ratio C2H6/O2/N2 = 9/7/84: (a) Temperature profiles; (b) Conversion and yield

46
profiles at Tb=400 ºC; (c) Conversion and yield profiles at Tb=440 ºC; (d) Conversion and

yield profiles at Tb=480 ºC (For (b)-(d): ( ) C2H6 conversion; ( ) O2 conversion; ( )

C2H4 yield; ( ) H2O yield; ( ) CO yield and ( ) CO2 yield).

Figure 6. Reactor model predictions when ethane feed concentration is increased at

different % mole in a ramp change with a Tb = 440 ºC: (a) Temperature profiles; (b)

Conversion and yield profiles at 1 % mole of ethane; (c) Conversion and yield profiles at 9

% mole of ethane; (d) Conversion and yield profiles at 18 % mole of ethane; (e)

Conversion and yield profiles at 40 % mole of ethane (For (b)-(e): ( ) C2H6 conversion; (

) O2 conversion; ( ) C2H4 yield; ( ) H2O yield; ( ) CO yield and ( ) CO2 yield).

Figure 7. Fractional surface coverage along the reactor at differ bath temperatures with an

inlet molar ratio C2H6/O2/N2 = 9/7/84.

47
100

80
Selectivity [%]

C2H4 Kinetic model


60
CO
CO2

40

20

0
0 20 40 60 80 100
Ethane conversion [%]

Fig. 1

48
15

50
10
Fn calculated [mmol/h]

40 5

0
30 0 5 10 15

C2H6
O2
20
C2H4
CO
10 CO2
H2O

0
0 10 20 30 40 50
Fn observed [mmol/h]

Fig. 2

49
1.00
Fractional coverage [-] (a)
0.95

θ∗
0.90
θΟ2
0.10 θΗ2Ο

0.05

0.00
400 420 440 460 480
Temperature [ºC]

1.0
(b)
0.8
Fractional coverage [-]

0.6 θ∗
θΟ2
θΗ2Ο
0.4

0.2

0.0
0 10 20 30 40 50 60 70
Wcat/Fethane,o [gcat h/molethane,o]

Fig. 3

50
550 (a)
500
Temperature [ºC]

450

400

Experimental observation
350 HTP estimated with BLA
HTP estimated with CA
300
0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m]

460
(b)
450
Temperature [ºC]

440

430
HTP estimated with BLA
HTP estimated with CA

420
0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m]

Fig. 4

51
560
1.0
(a) (b)
520

1- Conversion and Yield


0.9
Temperature [ºC]

Tb=480 ºC
480
0.8 - - - - Initial steady state
Tb=440 ºC - - - - Transient state
0.2
440

Tb=400 ºC
400 0.1
Initial steady state Second steady state
Final steady state Transient state
360 0.0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m] Reactor length [m]

1.0
1.0
0.9 (c)
(d) - - - - Final steady state
1- Conversion and Yield

- - - - Transient state
1- Conversion and Yield 0.8
0.8

0.7 0.6
- - - - Second steady state
- - - - Transient state
0.3
0.4
0.2
0.2
0.1

0.0 0.0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m] Reactor length [m]

Fig. 5

52
480 1.0
Initial steady state
(a) Second steady state (b)
470 Third steady state

1- Conversion and Yield


Final steady state 0.8
Transient state
Temperature [ºC]

460
40 mole % of ethane 0.6
18 mole % of ethane - - - - Initial steady state
450 - - - - Transient state
9 mole % of ethane
0.4
440
1 mole % of ethane
0.2
430

420 0.0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m] Reactor length [m]

1.0
1.0
(c)
(d)
1- Conversion and Yield

0.8
1- Conversion and Yield 0.8

0.6
- - - - Second steady state
0.6
- - - - Transient state - - - - Third steady state
- - - - Transient state
0.4 0.4

0.2 0.2

0.0 0.0
0.0 0.5 1.0 1.5 2.0 2.5 0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m] Reactor length [m]

1.0
(e)
1- Conversion and Yield

0.8

0.6
- - - - Final steady state
- - - - Transient state
0.4

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m]

53
1.0

0.8
Fractional coverage [-]

0.6 θΗ2Ο Tb=400 ºC


θ∗ Tb=440 ºC
θΟ2 Tb=480 ºC
0.4

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5
Reactor length [m]

Fig. 7

54
Table 1. Range of experimental conditions during the steady state in the ODH-Et

experiments.

Operational condition Unit Range

pC H kPa 5 – 22
2 6 ,o

pC H kPa 5 – 22
2 4 ,o

pO kPa 5 – 22
2 ,o

γ C 2 H 6 = y O 2 ,o y −C12 H 6 ,o molO 2 ,o molC−12 H6 ,o 0.5 – 2.0

γ C 2 H 4 = y O 2 ,o y C−12 H 4 ,o molO 2 ,o molC−12 H 4 ,o 0.5 – 2.0

Wcat FC−21H 6 ,o
g cat h mol−C12 H 6 ,o 10 - 140

Wcat FC−21H 4 ,o g cat h mol−C12 H 4 ,o 10 - 130

T ºC 400 - 480

X C2H6 molC 2H 6 mol−C12 H 6 ,o 0.17 - 0.85

X C2H 4 molC 2H 4 molC−12 H 4 ,o 0.03 - 0.14

55
Table 2. Reaction steps and catalytic cycles considered for the kinetic model to describe

the ODH-Et.

Step Elementary reaction step σI σII σIII σIV σV

A O2 + 2S 2O-S 1 7 5 3 2

1 C2H6 + O-S → C2H4 + H2O-S 2 0 0 0 0

2 C2H6 + 7O-S → 2CO2 + 3H2O-S + 4S 0 2 0 0 0

3 C2H6 + 5O-S → 2CO + 3H2O-S + 2S 0 0 2 0 0

4 C2H4 + 6O-S → 2CO2 + 2H2O-S + 4S 0 0 0 1 0

5 C2H4 + 4O-S → 2CO + 2H2O-S + 2S 0 0 0 0 1

B H2O-S H2O + S 2 6 6 2 2

56
Table 3. Reactor dimensions, catalyst dimensions, parameters and operational conditions

used in the reactor model.

Value Reference
Reactor and catalyst dimensions
L, mr 2.5 -
dt, mr 0.0256 [45]
dp, ms 0.0082 [45]
ε, m3f m-3r 0.48 [68]
ρb, kgcat m-3 r 75 -
Operational conditions
P, atm 1 -
Tb, ºC 400 - 480 -
To, ºC 200 -
Rep 1400a -
Flow rate, N m3 h-1 4 [68]
Ethane feed concentration, % mole 1 - 40 -
Oxygen feed concentration, % mole 7 -
Transport parameters
Deffr, mr2 h-1 32 [68,69]
2 -1
Deffz, mr h 53 [68,69]
3 -2 -1
kg, mf ms h 576 [68,69]
hg, kJ ms-2 (h K)-1 928.8 [68,69]
keffr, kJ (mr h K)-1 8.35b [45]
keffr, kJ (mr h K)-1 9.72c [45]
hw, kJ mr-2 (h K)-1 468b [45]
-2 -1 c
hw, kJ mr (h K) 1051.2 [45]
a b c
calculated at standard conditions, estimated by CA, estimated by BLA

57
Table 4. Kinetic parameters values and corresponding 95% probability confidence intervals

of the kinetic model used to describe the ODH-Et over the MoVTeNbO catalyst.

Parameter Estimated value Lower limit Upper limit t-value

A '1 , mmol (gcat h)-1 5.50E+00 5.49E+00 5.52E+00 7.59E+02

A '2 , mmol (gcat h)-1 6.86E-01 6.84E-01 6.88E-01 6.79E+02

A '3 , mmol (gcat h)-1 1.58E+00 1.57E+00 1.58E+00 6.61E+02

A '4 , mmol (gcat h)-1 2.60E+00 2.59E+00 2.61E+00 4.80E+02

A '5 , mmol (gcat h)-1 7.87E-01 7.84E-01 7.90E-01 5.14E+02

E A ,1 , kJ mol-1 9.05E+01 9.02E+01 9.07E+01 7.90E+02

E A ,2 , kJ mol-1 1.65E+02 1.64E+02 1.66E+02 5.43E+02

E A ,3 , kJ mol-1 1.50E+02 1.49E+02 1.50E+02 7.91E+02

E A ,4 , kJ mol-1 1.39E+02 1.38E+02 1.39E+02 5.31E+02

E A ,5 , kJ mol-1 1.32E+02 1.31E+02 1.32E+02 4.96E+02

o
−∆SO2 , J (mol K)-1 2.15E+02 2.14E+02 2.16E+02 8.33E+02

o
−∆SH O , J (mol K)-1 4.20E+01 4.19E+01 4.21E+01 6.75E+02
2

o
−∆HO2 , kJ mol-1 4.56E+01 4.55E+01 4.57E+01 8.56E+02

o
−∆HH2O , kJ mol-1 1.28E+02 1.28E+02 1.28E+02 7.48E+02

m2 9.22E-01 9.19E-01 9.24E-01 7.37E+02

m3 9.06E-01 9.03E-01 9.08E-01 7.23E+02

58
m4 1.23E+00 1.22E+00 1.23E+00 5.79E+02

m5 9.05E-01 9.02E-01 9.08E-01 6.37E+02

Note: F value = 7642, Ftab = 2.79, ttab = 1.97 at 1-α = 0.95 and 396 degrees of freedom.

59
Highlights:
• This work simulates the industrial catalytic behavior of ODH-Et on MoVTeNbO
• Simulations are carried out using a 2D pseudo-heterogeneous model
• A kinetic model is developed and it describes successfully laboratory experiments
• Industrial reactor model accounts for the effect of hydrodynamics on heat transfer
• Wall-cooled packed-bed reactor is an ideal design for the ODH-Et on MoVTeNbO

60

You might also like