Next Article in Journal
Metal–Organic Framework-Derived Rare Earth Metal (Ce-N-C)-Based Catalyst for Oxygen Reduction Reactions in Dual-Chamber Microbial Fuel Cells
Previous Article in Journal
Enhanced Catalytic Synthesis of Flavonoid by UV-B Radiation in Artemisia argyi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Kinetic Model for Catalytic N-Butane Oxidative Dehydrogenation under Oxygen-Free Reaction Conditions in a Fluidized CREC Riser Simulator

by
Abdulhamid Bin Sulayman
and
Hugo de Lasa
*
Chemical Reactor Engineering Centre, Department of Chemical and Biochemical Engineering, University of Western Ontario, London, ON N6A 5B9, Canada
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(8), 505; https://doi.org/10.3390/catal14080505
Submission received: 6 June 2024 / Revised: 17 July 2024 / Accepted: 29 July 2024 / Published: 5 August 2024
(This article belongs to the Special Issue Catalyzing the Sustainable Process Paradigm)

Abstract

:
This study considers the development of a kinetic model for the n-butane oxidative dehydrogenation (ODH) to C4-olefins using a VOx/MgO−γAl2O3 catalyst. The prepared catalyst contained 5 wt% V on an MgO modified γAl2O3 support. The developed catalyst exhibited both weak and medium acid sites, as revealed by NH3-temperature-programmed desorption. TPR/TPO analyses also indicated that 73% of the loaded VOx was reducible. Kinetic experiments were conducted in a fluidized CREC Riser Simulator at temperatures ranging from 475–550 °C and residence times of 5–20 s. An optimal C4-olefin selectivity of 86% was achieved at 500 °C and 10 s, with this selectivity then decreasing at higher temperatures and longer residence times. The kinetic model developed involved a Langmuir–Hinshelwood-type of kinetics that incorporated cracking, oxydehydrogenation, and complete oxidation reactions. Model parameters were determined by fitting experimental data with kinetic parameters established with narrow 95% confidence intervals and low cross-correlation.

Graphical Abstract

1. Introduction

The production of C4-olefins through the oxidative dehydrogenation (ODH) of n-butane is a potentially important industrial process. C4-olefins are in high demand due to being used to make petrochemical products like plastics, rubber, and resins [1,2,3]. Traditionally, olefins have been produced through fluid catalytic cracking (FCC), steam cracking, and the catalytic dehydrogenation of light paraffins. In FCC, olefins are a side product, while in steam cracking and in the dehydrogenation of paraffins, olefins are obtained directly [4,5,6]. These conventional processes face the following significant challenges: (a) a significant amount of thermal energy is required due to the endothermic reactions involved, (b) product separation is costly, and (c) catalysts are deactivated by coke, leading to frequent plant shutdowns [7,8,9,10].
The ODH process offers, instead, a more efficient alternative to produce C3 and C4-olefins with high selectivity. The overall ODH reaction is exothermic and can therefore, be conducted at lower temperatures. In addition, coke formation is limited. Consequently, the life of the catalyst can be extended significantly [11,12]. However, achieving high olefin selectivity is challenging, given the non-negligible amounts of complete oxidation products such as carbon dioxide. Thus, ODH requires efficient catalysts that limit complete n-butane oxidation [13,14]. These catalysts should supply solid-phase oxygen without having oxygen in the gas phase (gas-phase oxygen-free conditions), boosting, in this manner, olefin selectivity. In this way, metal oxide-based catalysts can function both as catalysts and as solid-phase oxygen providers, with their lattice oxygen favoring ODH olefin selectivity [15,16].
The ODH that converts n-butane into C4-olefins is described in Figure 1. It involves a partial oxidation step where C4-olefins are produced. However, several phenomena compete with this desirable reaction: (a) complete oxidation, during which both n-butane and formed C4-olefins are converted into carbon oxides (CO2 and CO); (b) the formation of coke, which may lead to catalyst deactivation;(c) a combined ODH and cracking, during which small amounts of unwanted lighter hydrocarbons, such as CH4, C2H4,C2H6, C3H6 and C3H8, are formed [17,18,19]. One should mention that these unwanted lighter hydrocarbons and coke can be further oxidized to form carbon oxides.
Thus, successful ODH requires the development of effective catalysts to improve the desired ODH reaction while minimizing unwanted side effects, such as cracking, coke formation, and complete hydrocarbon species oxidation.
Vanadium oxide-based catalysts are among the metal/support ODH catalysts that have been examined in previous research [20,21,22,23]. Furthermore, a series of magnesium oxide-modified vanadium oxide-based catalysts for alkane ODH has also been recently considered by our research group. These catalysts were characterized using BET, TPD/TPR, MH3-TPD, XRD, FTIR, LRS, and XPS. In particular, XPS analysis showed that the dominant catalyst oxidation species on the MgO-γAl2O3 catalyst are likely a blend of the V+4 and V+5 species [24].
The kinetics of the oxidative dehydrogenation (ODH) of n-butane to produce olefins have only been studied in a few works. A modified Mars van Krevelen (MK) mechanism was proposed by Grabowski et al. (2006) [25]. This work, which was developed in a fixed bed reactor under anaerobic conditions, suggests that alkane oxidation occurs through a reaction with the lattice oxygen of the metal oxide catalysts. In this kinetic model, it is assumed that the lattice oxygen of the catalyst is responsible for olefin formation while the surface oxygen of the catalyst contributes to the complete oxidation of hydrocarbons into carbon oxides.
Rubio and colleagues (2003) proposed an updated redox model for n-butane ODH. The reaction experiments were performed under anaerobic conditions in a fluidized bed reactor. This model distinguishes between the strongly bounded lattice oxygen and the loosely bounded oxygen of the catalyst studied [26]. This research suggests that the strongly bounded lattice oxygen is more selective towards olefin formation and the loosely bounded oxygen is better for producing carbon oxides. These researchers also considered that ODH is influenced by the oxidation state of the catalyst, with the oxygen species readily reacting with the fed alkanes and the olefin produced contributing to the formation of COx products.
The possible implementation of ODH under gas-phase oxygen-free conditions in risers and downer reactors presents new and valuable technological opportunities. These advantages have been demonstrated by our research team for both ethane and propane ODH under gas-phase oxygen-free conditions in a mini-fluidized CREC Riser Simulator [27,28]. The data obtained in this study have led to phenomenological-based kinetics with estimated parameters displaying narrow 95% confidence intervals [15,27,28,29]. Given these encouraging results, and the high prospects of implementing ODH on the industrial scale, research for N-butane ODH employing a VOx support on an MgO-γAl2O3 catalyst, under gas-phase oxygen-free conditions, has been recently developed [24]. This study included an assessment of catalyst synthesis, catalyst characterization, and catalyst performance. Furthermore, based on these positive outcomes, a kinetics and reaction network is established in this article that involves the catalyst lattice oxygen and dominant product species, such as butane, butenes, COx and coke, as well as their changes with reaction time. The reported rate equations are highly valuable, given that they can assist in the ODH catalytic reactor development and process scale-up.

2. Results and Discussion

2.1. Catalyst Characterization

Table 1 reports catalyst characterization results. In this respect, the H2-TPR analysis of the 5 wt% V/MgO−γAl2O3 (1:1) BODH catalyst reveals a distinct peak in the 475–485 °C range, showing controlled oxygen release. In addition, H2-TPR displays a higher rate of hydrogen consumption. Thus, on this basis, one can consider that a greater amount of oxygen is available in the 5 wt% V/MgO−γAl2O3 (1:1). This explains the catalyst’s ability to assist in producing highly selective C4-olefins. A more detailed description of catalyst properties can be found in Bin Sulayman et al. [24].
Furthermore, results obtained from NH3-TPD indicate that the 5 wt.%V/γAl2O3 catalyst has a higher acidity than the 5 wt.% V/MgO−γAl2O3 with MgO. This reduction in acidity can be attributed to the role of the MgO in moderating alumina acidity as well as blocking the acid sites within γAl2O3.
The results in Table 1 show the changes on the γ-alumina surface area when adding vanadium and magnesium. In that case, the 208 m2/g for γ-alumina support decreases to 177 m2/g and 152 m2/g for added vanadium and added magnesium, respectively. This reduction in surface area is attributed to the blockage of micropores in the γ-alumina, as the vanadium and magnesium particles can fill or obstruct the tiny support pores. Furthermore, when vanadium and magnesium are added together, it is observed that the vanadium and magnesium phases help each other to achieve a better-added oxide dispersion, leading, as a result, to an average pore size similar to the one for γ-alumina, with mild differences in specific surface area.

2.2. Oxidative Dehydrogenation of N-Butane in the CREC Riser Simulator

Figure 2 illustrates the influence of both reaction time and temperature on N-butane conversion. The data reported are the average values for four of the six consecutive n-butane injections in the CREC Riser Simulator. Typical standard deviations for repeat runs were in the ±7.16–9.56% range, with carbon balances closing in the 95% range, as documented in the upcoming Table 2. The results obtained for the first and second injections were neglected from further analysis, given that they tended to yield lower olefin selectivities. This behaviour was assigned to the catalyst-surface-adsorbed oxygen, available for ODH, following catalyst regeneration. One should note, in this respect, that the stability of the average n-butane conversion values for the four successive n-butane injections allowed one to neglect catalyst deactivation, an important kinetic model assumption.
Figure 2 reports that both temperature and reaction time have positive effects on n-butane conversion, which rises from 9% at 475 °C in 5 s to 37% at 550 °C and 20 s. Thus, longer contact times and higher temperatures are desirable properties for promising ODH, given that these conditions lead to more effective lattice oxygen utilization.
Figure 3 shows the influence of temperature and reaction time on product selectivity. The data reported here correspond again to the average values for data collected from four of the six n-butane consecutive injections. In the data reported in Table 2, one can see that there is a narrow range of selectivity variations, which allows one to disregard catalyst deactivation effects during successive runs.
Figure 3 reports that longer reaction times lead to a decrease in C4-olefin selectivity while simultaneously increasing COx production. Thus, on this basis, shorter contact times (e.g., 10 s), which favorably facilitate ODH in downer or riser units, are preferred. Furthermore, in Figure 3, one can observe that 500 °C provides an optimum selectivity, reaching 90% butene selectivity, at 10 s. This behaviour can be assigned to the difference in magnitude between pre-exponential factors and energies of activation for ODH and COx formation. As a result, one can conclude that the kinetic constant values reflect the significant influence of temperature and contact time on reaction selectivity, favoring butene formation at specific optimal conditions.
Table 2 reports the various chemical species selectivities and butane conversions obtained for one complete run at 10 s reaction time and 500 °C, 525 °C, and 550 °C. Every run was developed using six consecutive injections, with catalyst regeneration being conducted after every run. All runs, as with the ones reported in Table 2, were repeated three times. One can observe that, for injections 3–6, reproducible butane conversions and C2, C3, C4 olefin selectivities were obtained. All this points to the adequate and stable performance of the 5 wt%V/MgO-γAl2O3 catalyst. It has to be mentioned that only injection 1 and 2 yielded some extra butane conversion, likely due to the butane reaction with the catalyst labile oxygen, and, as a result, they were not used in the kinetic model analysis.
One should note, as shown in Table 2, that carbon-containing products other than C4-olefins were also detected. These were CO, CO2, CH4, C2H4, C2H6, C3H6, and C3H8. In particular, the percentage yields of the CH4, C2H4, C2H6, C3H6, and C3H8 species were below 2% (Table 2), and, as a result, their corresponding formation and consumption rates, as shown in Figure 1, are disregarded in the upcoming kinetics analysis.
Furthermore, the possible influence of thermal cracking in the ODH kinetics was also deemed negligible. This was the case given that, under the most extreme conditions, thermal cracking (temperature: 550 °C, reaction time: 20 s) was found to contribute no more than 2% to n-butane conversion.
Finally, experiments developed in the CREC Riser Simulator also allowed one to ascertain that the C4-Olefin/H2 ratio significantly surpassed the value of 1 while falling consistently in the 3.09–3.99 range. This supports the hypothesized oxidative dehydrogenation of n-butane, via catalytic ODH. Further information about this matter can be found in a previous study [24].

2.3. Kinetic Parameters Estimation

The kinetics of ODH can be based on the reaction network described in Figure 1 using a number of sound simplifications, as described in Section 2.2, with an effectiveness factor of 1; this gives an estimated Thiele Modulus lower than 1.
On this basis, the rate of N-butane consumption can be expressed as follows:
dp C 4 H 10 dt = W c RT V R r 1 + r 2 = W c RT V R k 1 + k 2 K C 4 H 10 p C 4 H 10 1 + K C 4 H 10 p C 4 H 10 + K C 4 H 8 p C 4 H 8 + K COX p COX × exp λ X C 4 H 10
Furthermore, the rate of butene formation can be described as:
dp C 4 H 8 dt = W c RT V R r 1 r 3 = W c RT V R k 1 K C 4 H 10 p C 4 H 10 k 3 K C 4 H 8 p C 4 H 8 1 + K C 4 H 10 p C 4 H 10 + K C 4 H 8 p C 4 H 8 + K COX p COX × exp λ X C 4 H 10  
Finally, the rate of COX formation can be expressed as follows:
dp COX dt = W c RT V R 4 r 2 + 4 r 3 = 4 W c RT V R k 2 K C 4 H 10 p C 4 H 10 + k 3 K C 4 H 8 p C 4 H 8 1 + K C 4 H 10 p C 4 H 10 + K C 4 H 8 p C 4 H 8 + K COX p COX × exp λ X C 4 H 10  
with VR (cm3) representing the volume of the reactor, Wc (g) standing for the weight of the catalyst, pi (atm) denoting the partial pressure of the “i” species, R (cm3 atm mol−1K−1) being the universal gas constant, T (k) representing the reactor temperature, and t (sec) denoting the reaction time.
Regarding Equations (1) to (3), and due to the absence of activity decay after the second n-butane injection, the λ decay parameter was set to zero for the analyses of the third, fourth, fifth, and sixth injections.

2.4. Adsorption Constants

Equations (1)–(3) include adsorption constants. These parameters may display strong correlations, and, as a result, one must be careful when selecting the approach to be adopted for their numerical derivations.
To address this issue, adsorption studies were developed in the mini-fluidized CREC Riser Simulator, with adsorption constants for the various chemical species on the γ-alumina support being determined using the following equations:
V i A V m = K i p i 1 + K i p i  
K i = K i 0 exp Δ H i R 1 T 1 T m  
with “ V i A ” referring to the volume of the species that is adsorbed on the surface, Vm representing the volume of monolayer, Ki denoting the species adsorption constant (atm−1), pi standing for the species’ partial pressure (atm), K i 0 being the pre-exponential factor of the adsorption constant (atm), −ΔH representing the heat of adsorption (kJ/mole), and Tm, being the selected median temperature.
On this basis, the adsorption parameters were calculated using a regression fitting process conducted by using a Nonlinear ModelFit built function in Wolfram Mathematica [24]. Table 3 reports the adsorption constants and heats of adsorption for the dominant species involved in the ODH butane conversion. The narrow 95% confidence spans show the limited parameter variation and, as a result, their potential phenomenological value.

2.5. Intrinsic Kinetic Parameters

The estimation of the six parameters ( k 1 * 0 = k 1 0 K C 4 H 10 0 , k 2 * 0 = k 1 0 K C 4 H 10 0 ,   k 3 * 0 = k 3 0 K C 4 H 8 0 , E1, E2, and E3) was conducted using nonlinear least-squares regression. The MATLAB function “lsqnonlin” was employed for this regression analysis. The numerical integration of the differential system, described in Equations (1)–(3), and the determination of the 95% confidence intervals for each estimated parameter were performed using the MATLAB functions “ode45” and “nlparci”, respectively. It was assumed that all reaction rate constants and activation energies had to lead to positive values during optimization.
The six kinetic parameters were estimated using a Trust Region Reflective Method, which minimized the objective function, as described in Equation (6):
SSQ = i = 1 N p i , experimental p i , theoretical 2
With this end in mind, various partial pressures of “i” components (n-butane, butene and COx species) obtained both experimentally, and predicted via the kinetic model, were considered in parameter regression using Equation (6). The lowest sum of squares (SSQ), as per Equation (6), was used for determining correlation coefficients (R2) and achieving model discrimination.
Table 4 reports the six estimated parameters, along with their corresponding 95% confidence intervals (CIs), when using a 5 wt% V/MgO−γAl2O3 catalyst.
As reported in Table 4, a high DOF (degrees of freedom) was used in this analysis to adequately calculate the six model parameters. The degree of freedom (DOF) was 570 and accounted for 576 experimental data points, minus the six kinetic constants that were numerically regressed. This was the case, given that every experimental condition considered included the third to sixth butane consecutive injections (excluding the first and second ones), with runs being repeated three times. The catalyst was regenerated after every complete run.
Table 4 shows that the estimated parameters were defined with reduced and desirable 95% confidence intervals. The other important information that can be seen in Table 4 is the low degree of cross-correlation between parameters (smaller than 1), with all reported cross-correlation coefficients being below 0.89.
When analyzing the activation energies reported in Table 4, several noteworthy observations emerged: (a) firstly, the BODH (Equation (1)) exhibits a favorable butene selectivity, given the large 2.64 × 10−5 mol.gcat−1.s−1 frequency factor; (b) conversely, despite the relatively high activation energy of 42 kJmol−1, the transformation of N-butane to COx (Equations (1) and (3)) is constrained by a very low-frequency factor; (c) lastly, the Equation (2) reaction, which involves the COx obtained from the C4-olefins, is substantially limited due to the combination of a 44 kJmol−1 activation energy and a small 2.32 × 10−7 mol.gcat−1.s−1 frequency factor. These findings are consistent with n-butane ODH kinetics, which confirm a triangular reaction network.
Furthermore, the high reaction rate constants for the formation of C4-olefins from N-butane through ODH point toward a highly selective C4-olefins process. Additionally, the very low kinetic constant required for complete C4-olefin oxidation can be explained given the expected role of magnesium oxide. Magnesium oxide reduces surface acidity, which inhibits C4-olefin re-adsorption and thus contributes to high C4-olefin selectivity.
To gain further insight into the accuracy of the proposed kinetic model, and that of the estimated parameters, the reactant and product partial pressure model predictions were compared with the experimental data. Figure 4a–d illustrate these comparisons. It is noteworthy to mention that the model predictions align closely with the experimental data, thus validating the proposed BODH reaction model, within the limits of experimental error.
Furthermore, the analysis presented in Figure 5 shows that the data do not exhibit any clear clustering patterns, suggesting that there are no significant external factors influencing the observed conversion changes. Additionally, the cross-correlation matrix, as shown in Table 4, reveals that the estimated parameters are weakly correlated. This further validates the reliability of our proposed kinetic model. Overall, these findings demonstrate that both the established adsorption and the kinetic parameters, along with the developed model, predict BODH reaction rates within the range of the operating conditions studied.
In this study, we compared the activation energies of BODH, obtained using different supported VOx catalysts, with values reported in the technical literature. Our findings, presented in Table 5, demonstrate that the 5 wt%V/MgO−γAl2O3 fluidizable catalyst exhibits a similar activation energy to other catalysts used for the formation of C4-olefins from N-butane (Step 1). Additionally, our research indicates that the 5 wt%V/MgO−γAl2O3 catalyst has the lowest activation energy for C4-olefins oxidation compared to those reported in the literature (Equation (3)). Thus, these results confirm that the C4-olefins selectivity is enhanced by a severe constrained placed on C4-olefins in regard to being further C4-olefin oxidation into carbon oxides.
Furthermore, the 5 wt%V/MgO−γAl2O3 fluidizable catalyst that was developed exhibited minimal coke formation (<0.02 wt %), as demonstrated in Table 5. This is a remarkable feature. This finding supports the conclusion that the λ decay parameter can be considered negligible in this case, as highlighted in this study. This beneficial characteristic of the 5 wt%V/MgO−γAl2O3 catalyst is a significant feature that is not present in other vanadium-supported catalysts for BODH [12,30,33,34,35].
In summary, the catalytic experiments conducted in this study, using a 5 wt%V/MgO−γAl2O3 fluidizable catalyst, in the CREC Riser Simulator provide a strong corroboration of the value of a phenomenologically based kinetic model. This BODH kinetics has the significant advantage of facilitating the development of a BODH process that employs a twin circulating fluidized bed configuration consisting of a BODH reactor and a dense fluidized bed catalyst regenerator [15]. The integrated BODH fluidized bed process is projected to achieve a 26% conversion of n-butane with a valuable 87% selectivity towards C4-olefins. Thus, this makes the ODH a promising avenue to explore further in regard to the production of olefins and, in particular, butenes.

3. Experimental Methods

3.1. Catalyst Synthesis and Characterization

The catalyst employed in this study, consisting of a 5 wt%V/MgO−γAl2O3, was prepared using a wet saturation impregnation technique. The γAl2O3 support (SASOL, Sasolburg, South Africa, Catalox SSCa 5/200) underwent modification with magnesium, and the vanadium loading was achieved using an ammonium metavanadate (NH4VO3; Sigma–Aldrich, Saint Louis, MO, USA, 99%) precursor. Various characterization techniques, including BET surface area analysis, X-ray diffraction (XRD), H2-temperature-programmed reduction (H2-TPR), NH3-temperature-programmed desorption (TPD), pyridine Fourier transform infrared spectroscopy (FTIR), laser Raman spectroscopy (LRS), and X-ray photoelectron spectroscopy (XPS), were used to characterize the catalyst. The detailed syntheses and characterization procedures for the catalyst, used in oxidative dehydrogenation (ODH), can be found in a recent publication by Bin Sulayman et al. [24].

3.2. Oxidative Dehydrogenation of N-Butane Using Catalyst Lattice Oxygen

The BODH experiments were conducted by using a 5 wt% V/MgO−γAl2O3 catalys, in the CREC Riser Simulator, which is a bench-scale mini-fluidized bed reactor. The reactor operates under batch conditions and is designed to evaluate catalysts and conduct kinetics relevant for circulating fluidized beds, such is the case of risers and downers [36]. An overview of the operating procedures and system configurations was previously described by de Lasa [36]. Additional details regarding the CREC Riser Simulator unit are provided in Appendix A.
In the experiments of this study, six consecutive injections of butane were introduced into the CREC Riser Simulator without having the catalyst regenerated between injections. The runs were conducted at temperatures ranging from 475 °C to 550 °C and contact times varying from 5 to 20 s. Each run utilized 0.7 g of catalyst and 4 mL butane injections while maintaining a constant catalyst-to-feed weight ratio. The degree of catalyst reduction was monitored by comparing its initial oxygen content with its oxygen content after each injection. Instantaneous conversion rates and selectivities to primary products were calculated following every injection. To ensure result reproducibility, all thermal and catalytic experiments were repeated three times.
The analyses of reactor effluents and the methods used to calculate n-butane conversion and product selectivities were detailed in recent publications [24]. Carbon mass balances, accounting for all carbon-containing products (carbon monoxide, methane, carbon dioxide, ethylene, ethane, propylene, propane, and carbon deposited over the ODH catalyst), exceeded 95% in all cases [24].

4. Kinetic Modeling Methods

The oxidative dehydrogenation of n-butane involves a complex parallel-series reaction network. This reaction network can be simplified as follows: (a) the conversion of n-butane to C4-olefins, (b) the undesired formation of COx from n-butane, and (c) the secondary combustion of C4-olefins to COx. In these reactions, the catalyst lattice oxygen serves as the sole source of oxygen:
ODH of n-butane to C4 olefins:
  nC 4 H 10 + O VO X / MgO γ Al 2 O 3   C 4 Olefins   +   H 2 O  
Complete oxidation of n-butane:
nC 4 H 10 + 5 + 4 x O VO X / MgO γ Al 2 O 3   CO X + 5   H 2 O
Complete oxidation of C4 olefins:
C 4 olefins + 4 + 4 x O VO X / MgO γ Al 2 O 3   4   CO X + 4   H 2 O  
with −O− representing oxygen species on the vanadium oxide lattice.
Reactions as described via Equations (7) and (8) are considered as primary reactions, while a reaction as reported with Equation (9) is considered to be a secondary reaction step. These three reactions steps play a crucial role in determining the selectivity of C4-Olefins. Based on this, we propose a triangular parallel-series reaction network for the BODH process.
The proposed reaction network suggests that n-butane reacts with the lattice oxygen of the catalyst, leading to the production of both C4-olefins and COx. The associated rate constants for these reactions are denoted as k1 and k2, respectively. Additionally, C4-Olefins can undergo further combustion to form COx, with the associated rate constant being designated as k3.

4.1. Development of the Kinetic Model

In the n-butane oxidative dehydrogenation (BODH) study conducted and discussed in this work, in the absence of gas-phase oxygen, two types of oxygen species can be considered: (a) surface oxygen, or weakly adsorbed oxygen, and (b) lattice oxygen. The surface oxygen facilitates the combustion of n-butane and C4-Olefins to produce COx, while the lattice oxygen plays a key role in providing a controlled release of oxygen, which is crucial for high C4-Olefins formation.
According to the Langmuir–Hinshelwood mechanism, n-butane, C4-Olefins, and COx are adsorbed at equilibrium on a catalyst surface. Therefore, the surface reaction step is deemed to be the rate-controlling step, in contrast to the species adsorption and desorption steps. According to this mechanism, there are two types of sites: (a) Site1-[V2O5-O-], where lattice sites are in an oxidized state, (b) Site 1-[V2O5R], where there are reduced lattice oxygen sites and surface oxygen vacancies, and (c) Site 2-[S], which are the support sites.
Based on this, the following elementary steps can be considered as follows:
  • Adsorption of n-butane, on Site-2:
    C4H10 (g) + [S](s) ←⎯⎯→ C4H10 − [S](s).
  • Reaction between adsorbed n-butane in Site 2 and lattice oxygen in Site1-[V2O50] with C4-olefin formation:
    C4H10 − [S](s) + [V2O5-O] (s) → C4H8 − [S](s) + H2O(g) + [V2O5R](s).
  • Desorption of C4-olefin formed in Site 2:
    C4H8 − [S](s) ←⎯⎯→ C4H8 (g) + [S](s).
  • Reaction between adsorbed n-butane in Site 2 and lattice oxygen [V2O50] in Site 1 with formation of COx:
    C4H10 − [S](s) + (5 + 4x) [V2O5-O-](s) → 4COx(g) + 5H2O(g) + (5 + 4x) [V2O5R](s) + [S](s)
  • Reaction between adsorbed butene in Site 2 and lattice oxygen [V2O50] in Site 1 with formation of COx.
    C4H8 − [S](s) + (4 + 4x) [V2O5-O-] (s) → 4COx(g) + 4H2O(g) + (4 + 4x) [V2O5R](s) + [S](s).
  • Site 1 reoxidation with gas-phase oxygen during catalyst regeneration:
    [V2O5R] (s) +0.5 O2 (g) → [V2 O5-O-](s).
Thus, on this basis, and as described with Equations (7)–(9), it follows that:
r1 = k1θC4H10(1 − β)
r2 = k2θC4H10(1 − β)
r3 = k3θC4H8(1 − β)
where ki is the reaction rate constant (mol/gcat.s), ri stands as the reaction rate (mol/gcat.s), θi denotes the surface coverage of the adsorbed species “i”, and β represents the reduced vanadium site fractions. Furthermore, γ + β= 1 represents the reduced and oxidized vanadium site fractions, and θC4H10 + θC4H8 + θCOx + θv = 1 stand for the support-site fractions.
Therefore, the θi the support site fractions can be defined as follows:
θ C 4 H 10 = K C 4 H 10 C C 4 H 10 1 + K C 4 H 10 C C 4 H 10 + K C 4 H 8 C C 4 H 8 + K COX C COX  
θ C 4 H 8 = K C 4 H 8 C C 4 H 8 1 + K C 4 H 10 C C 4 H 10 + K C 4 H 8 C C 4 H 8 + K COX C COX  
with Ki representing the adsorption constant in cm3/mol and Ci denoting the concentration of species “i” in mol/cm3.
In order to account for the consumption of the lattice oxygen during BODH reactions, the proposed kinetic model incorporated a time-dependent oxidation extent variable. This extent variable was calculated by comparing the remaining oxygen content of the catalyst after the run to the initial oxygen content of the catalyst before the BODH run. As expected, the degree of catalyst oxidation decreased over successive reaction cycles, requiring reoxidation after a certain number of cycles. In our study, a total of six injections were introduced into the reactor without the reoxidation of the catalyst, given that, beyond this point, a decrease in C4-olefin selectivity and an increase in cracking products were observed. This change in selectivity can be attributed to the depletion of surface oxygen under these conditions. Thus, catalyst regeneration was performed after six cycles, corresponding to 1/6 of the total catalyst mass flow in a continuous n-butane ODH unit.
To represent the φ catalytic activity, a decay function based on the converted n-butane was reported to be suitable for gas-phase free oxygen BODH [15], as follows:
φ = exp λ X C 4 H 10  
where φ is the degree of oxidation of the catalyst, X is the n-butane conversion, and λ is a decay constant. This function has the benefit of accounting for the effects of reaction parameters (concentration, temperature, and reaction time) based on the extent of the oxidation of the catalyst.
According to Equations (16)–(18), β can be linked to φ as:
1 β = φ  
A substitution of Equations (19)–(22) into (16)–(18), can be used to express reaction rates in terms of the partial pressures of different species, as follows:
r 1 = k 1 K C 4 H 10 P C 4 H 10 1 + K C 4 H 10 P C 4 H 10 + K C 4 H 8 P C 4 H 8 + K COX P COX × exp λ X C 4 H 10    
r 2 = k 2 K C 4 H 10 P C 4 H 10 1 + K C 4 H 10 P C 4 H 10 + K C 4 H 8 P C 4 H 8 + K COX P COX × exp λ X C 4 H 10    
r 3 = k 3 K C 4 H 8 P C 4 H 8 1 + K C 4 H 10 P C 4 H 10 + K C 4 H 8 P C 4 H 8 + K COX P COX × exp λ X C 4 H 10  
This was the case, given the fact that the effects of coke formation on BODH were negligible (<0.02 wt%), thereby excluding any possibility that they could cause catalyst deactivation.

4.2. Kinetic Modeling in the Riser Simulator Reactor

BODH reactions are performed in a well-mixed mini-fluidized batch reactor designated as the CREC Riser Simulator [37]. Equation (26) can be used to describe the rates at which these reactions have been studied at the Chemical Reactor Engineering Centre labs at Western University, as follows:
η r i = V R W C   d p i RT dt  
with VR being the volume of the reactor (cm3), Wc standing for the catalyst weight (g), pi representing the partial pressure of a species “i” (atm), R denoting the universal gas constant, T being the temperature of the reactor (K), t standing for the reaction time, and η being the catalyst effectiveness factor.
In the case of an effectiveness factor of 1, the following Equation (27) can be used to determine the reaction rate:
r i = V R W C d p i RT dt  
As a result, the general rate of reaction for each chemical species can be calculated as follows:
dp i dt = W c RT V R r i  
Equations (23)–(25) and (28) can be further employed to derive Equations (1)–(3), already reported in Section 2.3 of the present study.

5. Conclusions

(a)
A reaction kinetics for n-butane ODH, using a 5 wt%V/MgO−γAl2O3 (1:1) fluidizable catalyst within an oxygen-free atmosphere, can be successfully established using data obtained from a fluidized bed CREC Riser Simulator.
(b)
The adsorption and intrinsic kinetic parameters from the postulated Langmuir–Hinshelwood rate model can be determined separately via independent adsorption and reaction experiments.
(c)
The evaluated intrinsic kinetic parameters for n-butane ODH (k10, k20, k30, E1, E2, and E3) can be successfully determined using a large degree of freedom (DOF) data set, with satisfactorily reduced 95% confidence intervals and low kinetic parameter cross-correlation.
(d)
The derived n-butane ODH kinetics provides a good estimation of n-butane conversions and butene selectivities at various temperatures and contact times. In particular, the postulated kinetics accurately predicts the n-butane conversions, ranging from 25% to 27%, and C4-olefins selectivities, ranging from 84% to 87%, at 10 s with negligible coke formation.
(e)
The derived n-butane ODH kinetics provides an excellent tool for riser and downer reactor industrial scale simulations, given that it was derived under operating conditions (temperature, reaction time, partial pressures, catalyst/n-butane weight ratios) close to the ones anticipated for large scale riser and downer units.

Author Contributions

A.B.S.: catalyst synthesis, catalyst characterization, and experimental runs in the CREC Riser Simulator, preparation of the manuscript. H.d.L.: catalyst development, data interpretation, discussion of results, review of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received funding from the Libyan Ministry of Higher Education and Scientific Research, via a scholarship awarded to Abdulhamid Bin Sulayman, and the Natural Science and Engineering Research Council of Canada (NSREC), via a NSERC Discovery Grant awarded to Hugo de Lasa.

Data Availability Statement

There are no additional data to be reported.

Acknowledgments

We would like to thank Florencia de Lasa, who assisted with the editing of this paper and the drafting of the graphical abstract.

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

COxCarbon oxides
EdesActivation energy of desorption (kJ/mol)
EiActivation energy for “i” species (kJ/mol)
kdesPre-exponential desorption factor (cm3/gcat.min)
kiReaction rate constant for “i” species (mol/gcat.s)
ki0Intrinsic kinetics constant pre-exponential factor for “i” species (mol/gcat.s)
KiAdsorption constant for for “i” species (atm−1)
Ki0Adsorption constant pre-exponential factor for “i” species (atm−1)
piPartial pressure of species “i” (atm)
ri“i” species reaction rate (mol/gcat.s)
RUniversal gas constant
SBETBrunauer−Emmet−Teller specific surface area (m2/g)
SiSelectivity for for “i” species (%)
TmMedian temperature in the 475–550 °C range (K)
VmVolume of monolayer coverage (cm3/gcat)
ViASpecies volume adsorbed on the catalyst (cm3/gcat)
V2O5vanadium oxide species at the reduced state
V2O5-Ovanadium oxide species at the oxidized state
XC4H10N-butane conversion (%)
YC4H8C4-olefins yield (%)
Greek Symbols
βDegree of reduction of catalyst
λDecay constant (−)
θSurface coverage of adsorbed species
φDegree of oxidation of the catalyst
ΔHiHeat of adsorption (kJ/mol)
Abbreviations
CRECChemical Reactor Engineering Centre
FIDFlame Ionization Detector
FTIRFourier Transform Infrared Spectroscopy
LRSLaser Raman Spectroscopy
BODHN-butane Oxidative Dehydrogenation
TCDThermal Conductivity Detector
TPDTemperature-Programmed Desorption
TPOTemperature-Programmed Oxidation
TPRTemperature-Programmed Reduction
XPSX-ray Photoelectron Spectroscopy
XRDX-ray Diffraction

Appendix A

Appendix A.1. CREC Riser Simulator—Equipment Description

The experimental runs for this study were performed in a 50 mL mini-fluidized bed reactor, designated as the CREC (Chemical Reactor Engineering Center) Riser Simulator. The CREC Riser Simulator reproduces, at the lab-scale, the conditions of an FCC riser and downer industrial units in terms of temperatures, contact times, hydrocarbon partial pressures, and C/O ratios. Thus, the data obtained from the CREC Riser Simulator unit experiments provide the necessary information to evaluate catalyst performance and develop kinetic models.
Figure A1 shows a CREC Riser Simulator with an impeller in its upper section that rotates generating a vortex and as a result, particle fluidization with high hydrocarbon recirculation in the catalyst basket.
Figure A1. Sectional view of the CREC Riser Simulator with details of the impeller and catalyst basket. The green arrows indicate the gas flow induced by the impeller [36].
Figure A1. Sectional view of the CREC Riser Simulator with details of the impeller and catalyst basket. The green arrows indicate the gas flow induced by the impeller [36].
Catalysts 14 00505 g0a1
The CREC Riser Simulator operates jointly with a vacuum box, a Mandel/Shimadzu GC2010 gas chromatograph, a series of sampling valves, a timer, two pressure transducers, and two temperature controllers. Figure A2 reports a schematic diagram of the reactor with its various accessories.
Once the reactor is set to the desired reaction temperature (e.g., 500 °C) and the catalyst is conditioned as required (e.g., regenerated after six consecutive injections of n-butane fed into the CREC Riser Simulator), runs are conducted as follows (refer to Figure A3): (A) manual injection of the n-butane feed, (B) catalytic reaction, (C) evacuation of the reactor contents to a vacuum box; all these are conducted using a four-port valve (4PV) and a timer once the desired reaction time (e.g., 10 s) is reached.
Figure A2. Schematic diagram of the CREC Riser Simulator and its associated system of auxiliary valves, vacuum box and GC equipment Notes: (a) Between A to B in Figure A3 (black broken lines) with “2” and “3” connected via the 4PV and, with “6” and “7” connected via the 6PV, (b) Between B to C in Figure A3 (full red lines) with 3 and 4 connected via the 4PV and with 4 and 5 lines connected via the 6PV [36].
Figure A2. Schematic diagram of the CREC Riser Simulator and its associated system of auxiliary valves, vacuum box and GC equipment Notes: (a) Between A to B in Figure A3 (black broken lines) with “2” and “3” connected via the 4PV and, with “6” and “7” connected via the 6PV, (b) Between B to C in Figure A3 (full red lines) with 3 and 4 connected via the 4PV and with 4 and 5 lines connected via the 6PV [36].
Catalysts 14 00505 g0a2
Figure A3. Pressure profile (blue line) for a 10 s reaction run in the CREC Riser Simulator following a 4 mL n-butane injection at 500 °C (A) manual injection of the n-butane feed, (B) catalytic reaction period, (C) evacuation of the reactor contents to a vacuum box. Pressure profile (red line) in the vacuum box. All the reported steps are performed using a four-port valve (4PV) and a timer once the desired reaction time (e.g., 10 s) is reached.
Figure A3. Pressure profile (blue line) for a 10 s reaction run in the CREC Riser Simulator following a 4 mL n-butane injection at 500 °C (A) manual injection of the n-butane feed, (B) catalytic reaction period, (C) evacuation of the reactor contents to a vacuum box. Pressure profile (red line) in the vacuum box. All the reported steps are performed using a four-port valve (4PV) and a timer once the desired reaction time (e.g., 10 s) is reached.
Catalysts 14 00505 g0a3

Appendix A.2. Experimental Procedures in the CREC Riser Simulator

For each catalytic experiment, the required amount of catalyst (0.7 g) was first loaded in the reactor basket, and then the reactor was closed. A temperature program was run to heat the reactor system to the desired reaction temperature. An argon flow at 20 mL/min was maintained during the reactor heating period to ensure that the reactor system was free of oxygen (air). Once the reactor reached the desired temperature, the argon flow was arrested and the pressure in the vacuum box was brought to 3 psi using a vacuum pump.
At this point, the impeller was turned on, and when it reached 5000 rpm, the feed (n-butane) was injected into the reactor using a preloaded syringe. During the reaction period, when the BODH reaction was taking place, the pressure profile of the reactor was monitored using a pressure transducer, as described in Figure A3. At the end of the pre-specified reaction time (e.g., 10 s), a valve isolating the reactor from a 1000 mL stainless steel vacuum-operated container, was opened and the contents of the CREC Riser Simulator were transferred from the reactor to the stainless-steel vacuum container. At this time, there was an abrupt decrease in the reactor pressure, as shown in Figure A3, with reactant and product species removed from the reactor almost instantaneously.
Finally, the product species were analyzed using a gas chromatograph. After six successive n-butane injections following the same procedure, the catalyst was regenerated (oxidized) by flowing air, at a specified temperature and time (e.g., 575 °C, 20 min), and prepared for the next run cycle.

References

  1. Gambo, Y.; Adamu, S.; Tanimu, G.; Abdullahi, I.M.; Lucky, R.A.; Ba-Shammakh, M.S.; Hossain, M.M. CO2-mediated oxidative dehydrogenation of light alkanes to olefins: Advances and perspectives in catalyst design and process improvement. Appl. Catalsis A Gen. 2021, 623, 118273. [Google Scholar] [CrossRef]
  2. Gao, Y.; Wang, X.; Corolla, N.; Eldred, T.; Bose, A.; Gao, W.; Li, F. Alkali metal halide–coated perovskite redox catalysts for anaerobic oxidative dehydrogenation of n-butane. Sci. Adv. 2022, 8, eabo7343. [Google Scholar] [CrossRef] [PubMed]
  3. Ronda-Lloret, M.; Slot, T.K.; van Leest, N.P.; de Bruin, B.; Sloof, W.G.; Batyrev, E.; Sepúlveda-Escribano, A.; Ramos-Fernandez, E.V.; Rothenberg, G.; Shiju, N.R. The Role of vacancies in a Ti2CTx MXene-derived catalyst for butane oxidative dehydrogenation. ChemCatChem 2022, 14, 7. [Google Scholar] [CrossRef]
  4. Gholami, Z.; Gholami, F.; Tišler, Z.; Tomas, M.; Vakili, M. A review on production of light olefins via fluid catalytic cracking. Energies 2021, 14, 1089. [Google Scholar] [CrossRef]
  5. Akah, A.; Williams, J.; Ghrami, M. An Overview of light olefins production via steam enhanced catalytic cracking. Cataysisl Surv. Asia 2019, 23, 265–276. [Google Scholar] [CrossRef]
  6. Zhang, X.; Gong, J.; Wei, X.; Liu, L. Increased light olefin production by sequential dehydrogenation and cracking reactions. Catalysts 2022, 12, 1457. [Google Scholar] [CrossRef]
  7. Luongo, G.; Donat, F.; Bork, A.H.; Willinger, E.; Landuyt, A.; Müller, C.R. Highly selective oxidative dehydrogenation of ethane to ethylene via chemical looping with oxygen uncoupling through structural engineering of the oxygen carrier. Adv. Energy Mater. 2022, 12, 2200405. [Google Scholar] [CrossRef]
  8. Zhou, J.; Zhao, J.; Zhang, J.; Zhang, T.; Ye, M.; Liu, Z. Regeneration of catalysts deactivated by coke deposition: A review. Chin. J. Catal. 2020, 41, 1048–1061. [Google Scholar] [CrossRef]
  9. Neal, L.M.; Haribal, V.P.; Li, F. Intensified ethylene production via chemical looping through an exergetically efficient redox scheme. iScience 2019, 19, 894–904. [Google Scholar] [CrossRef]
  10. Watanabe, R.; Suganuma, H.; Yoda, Y.; Karasawa, F.; Verma, P.; Fukuhara, C. Deactivation of an Fe-based catalyst in the dehydrogenation of light alkanes under H2S co-feeding: A case study. Appl. Catal. A Gen. 2024, 683, 119848. [Google Scholar] [CrossRef]
  11. Gambo, Y.; Adamu, S.; Abdulrasheed, A.A.; Lucky, R.A.; Ba-Shammakh, M.S.; Hossain, M.M. Catalyst design and tuning for oxidative dehydrogenation of propane—A review. Appl. Catal. A Gen. 2021, 609, 117914. [Google Scholar] [CrossRef]
  12. Khan, M.Y.; Adamu, S.; Lucky, R.A.; Razzak, S.A.; Hossain, M.M.; Hossain, M.M. Oxidative dehydrogenation of n-Butane to C4 Olefins using lattice oxygen of VOx/Ce-meso-Al2O3 under gas-phase oxygen-free conditions. Energy Fuels 2020, 34, 7410–7421. [Google Scholar] [CrossRef]
  13. Zhang, Z.; Tian, J.; Wu, X.; Surin, I.; Pérez-Ramírez, J.; Hemberger, P.; Bodi, A. Unraveling radical and oxygenate routes in the oxidative dehydrogenation of propane over boron nitride. J. Am. Chem. Soc. 2023, 145, 7910–7917. [Google Scholar] [CrossRef]
  14. Yan, B.; Li, W.C.; Lu, A.H. Metal-free silicon boride catalyst for oxidative dehydrogenation of light alkanes to olefins with high selectivity and stability. J. Catal. 2019, 369, 296–301. [Google Scholar] [CrossRef]
  15. Rostom, S.; de Lasa, H. Propane oxidative dehydrogenation on vanadium-based catalysts under oxygen-free atmospheres. Catalysts 2020, 10, 418. [Google Scholar] [CrossRef]
  16. Jiang, X.; Zhang, X.; Purdy, S.C.; He, Y.; Huang, Z.; You, R.; Wei, Z.; Meyer, H.M.; Yang, J.; Pan, Y.; et al. Multiple promotional effects of vanadium oxide on boron nitride for oxidative dehydrogenation of propane. JACS Au 2022, 2, 1096–1104. [Google Scholar] [CrossRef]
  17. McDermott, W.; Venegas, J.; Hermans, I. Selective oxidative cracking of n-Butane to light olefins over hexagonal boron nitride with limited formation of COx. ChemSusChem 2019, 13, 152–158. [Google Scholar] [CrossRef]
  18. Huš, M.; Kopač, D.; Bajec, D.; Likozar, B. Effect of surface oxidation on oxidative propane dehydrogenation over chromia: An Ab initio multiscale kinetic study. ACS Catal. 2021, 11, 11233–11247. [Google Scholar] [CrossRef] [PubMed]
  19. Kopač, D.; Jurković, D.L.; Likozar, B.; Huš, M. First-principles-based multiscale modelling of nonoxidative butane dehydrogenation on Cr2O3(0001). ACS Catal. 2020, 10, 14732–14746. [Google Scholar] [CrossRef]
  20. Kazerooni, H.; Towfighi Darian, J.; Mortazavi, Y.; Khadadadi, A.A.; Asadi, R. Titania-supported vanadium oxide synthesis by atomic layer deposition and Its application for low-temperature oxidative dehydrogenation of propane. Catal. Lett. 2020, 150, 2807–2822. [Google Scholar] [CrossRef]
  21. Xiong, C.; Chen, S.; Yang, P.; Zha, S.; Zhao, Z.J.; Gong, J. Structure-performance relationships for propane dehydrogenation over aluminum supported vanadium oxide. ACS Catal. 2019, 9, 5816–5827. [Google Scholar] [CrossRef]
  22. Liu, Q.; Yang, Z.; Luo, M.; Zhao, Z.; Wang, J.; Xie, Z.; Guo, L. Vanadium-containing dendritic mesoporous silica nanoparticles: Multifunctional catalysts for the oxidative and non-oxidative dehydrogenation of propane to propylene. Microporous Mesoporous Mater. 2019, 282, 133–145. [Google Scholar] [CrossRef]
  23. Tanimu, G.; Aitani, A.M.; Asaoka, S.; Alasiri, H. Oxidative dehydrogenation of n-butane to butadiene catalyzed by new mesoporous mixed oxides NiO-(beta-Bi2O3)-Bi2SiO5/SBA-15 system. Mol. Catal. 2020, 488, 110893. [Google Scholar] [CrossRef]
  24. Bin Sulayman, A.; Torres Brauer, N.; de Lasa, H. A Fluidizable catalyst for n-butane oxidative dehydrogenation under oxygen-free reaction conditions. Catalysts 2023, 13, 1462. [Google Scholar] [CrossRef]
  25. Grabowski, R. Kinetics of oxidative dehydrogenation of C2-C3 alkanes on oxide catalysts. Catal. Rev. Sci. Eng. 2006, 48, 199–268. [Google Scholar] [CrossRef]
  26. Cortés, I.; Rubio, O.; Herguido, J.; Menéndez, M. Kinetics under dynamic conditions of the oxidative dehydrogenation of butane with doped V/MgO. Catal. Today 2004, 91, 281–284. [Google Scholar] [CrossRef]
  27. Rostom, S.; de Lasa, H. High propylene selectivity via propane oxidative dehydrogenation using a novel fluidizable catalyst: Kinetic modeling. Ind. Eng. Chem. Res. 2018, 57, 45. [Google Scholar] [CrossRef]
  28. Al-Ghamdi, S.A.; Hossain, M.M.; de Lasa, H. Kinetic modeling of ethane oxidative dehydrogenation over VOx/Al2O3 catalyst in a fluidized-bed riser simulator. Ind. Eng. Chem. Res. 2013, 52, 5235–5244. [Google Scholar] [CrossRef]
  29. Elbadawi, A.A.H.; Ba-Shammakh, M.S.; Al-Ghamdi, S.; Razzak, S.A.; Hossain, M.M.; de Lasa, H. Phenomenologically based kinetics of ODH of ethane to ethylene using lattice oxygen of VOx/Al2O3–ZrO2 catalyst. Chem. Eng. Res. Des. 2017, 117, 733–745. [Google Scholar] [CrossRef]
  30. Lucky, R.A.; Adamu, S.; Khan, M.Y.; Razzak, S.A.; Hossain, M.M. Kinetics of oxidative dehydrogenation of n-butane to C4-Olefins over a VOx/CeO2-γAl2O3 catalyst in gas-phase oxygen-free Conditions. Ind. Eng. Chem. Res. 2020, 59, 17815–17827. [Google Scholar] [CrossRef]
  31. Dejoz, A.; López Nieto, J.M.; Melo, F.; Vázquez, I. Kinetic study of the oxidation of n-butane on vanadium oxide supported on Al/Mg mixed oxide. Ind. Eng. Chem. Res. 1997, 36, 2588–2596. [Google Scholar] [CrossRef]
  32. Madaan, N.; Haufe, R.; Shiju, N.R.; Rothenberg, G. Oxidative dehydrogenation of n-butane: Activity and kinetics over VOx/Al2O3 Catalysts. Top. Catal. 2014, 57, 1400–1406. [Google Scholar] [CrossRef]
  33. Sánchez-García, J.L.; Handy, B.E.; Ávila-Hernández, I.N.; Rodríguez, A.G.; García-Alamilla, R.; Cardenas-Galindo, M.G. Structure, acidity, and redox aspects of VOx/ZrO2/SiO2 catalysts for the n-butane oxidative dehydrogenation. Catalysts 2020, 10, 550. [Google Scholar] [CrossRef]
  34. Otroshchenko, T.; Jiang, G.; Kondratenko, V.A.; Rodemerck, U.; Kondratenko, E.V. Current status and perspectives in oxidative, non-oxidative and CO2-mediated dehydrogenation of propane and isobutane over metal oxide catalysts. Chem. Soc. Rev. 2021, 50, 473–527. [Google Scholar] [CrossRef]
  35. Hossain, M.M. Kinetics of oxidative dehydrogenation of propane to propylene using lattice oxygen of VOx/CaO/γAl2O3 Catalysts. Ind. Eng. Chem. Res. 2017, 56, 4309–4318. [Google Scholar] [CrossRef]
  36. de Lasa, H. Multifunctional Riser and Downer Simulator. USA Patent 10,220,363, 3 May 2019. [Google Scholar]
  37. de Lasa, H. The CREC fluidized riser simulator a unique tool for catalytic process development. Catalysts 2022, 12, 888. [Google Scholar] [CrossRef]
Figure 1. Catalytic ODH reaction network for C4H8 butene formation.
Figure 1. Catalytic ODH reaction network for C4H8 butene formation.
Catalysts 14 00505 g001
Figure 2. N-Butane conversion values at various reaction times and temperatures. The reported data are average values for four consecutive injections. Note: the typical standard deviation for repeat runs was ±7.16–9.56%.
Figure 2. N-Butane conversion values at various reaction times and temperatures. The reported data are average values for four consecutive injections. Note: the typical standard deviation for repeat runs was ±7.16–9.56%.
Catalysts 14 00505 g002
Figure 3. C4-Olefin and COx selectivities at various reaction times and temperatures. The reported data represent the average values for four consecutive n-butane injections. Note: typical standard deviations for repeat runs are ±5.6–6.27.
Figure 3. C4-Olefin and COx selectivities at various reaction times and temperatures. The reported data represent the average values for four consecutive n-butane injections. Note: typical standard deviations for repeat runs are ±5.6–6.27.
Catalysts 14 00505 g003
Figure 4. Comparison between the average experimental data and kinetic model predictions at 475 °C, 500 °C, 525 °C, and 550 °C. Typical standard deviation for three repeat runs: ±6.05–8.15%”. Notes: (a) The average experimental data points ( 1 3 p i , av 3 )   are   for   three repeat runs, with p i , av being calculated for four out of six consecutive injections (excluding the first and second one); (b) the ODH catalyst was regenerated after every sixth repeated injection runs; (c) the “i” subscript refers to the n-butane, butene, and Coax species under consideration.
Figure 4. Comparison between the average experimental data and kinetic model predictions at 475 °C, 500 °C, 525 °C, and 550 °C. Typical standard deviation for three repeat runs: ±6.05–8.15%”. Notes: (a) The average experimental data points ( 1 3 p i , av 3 )   are   for   three repeat runs, with p i , av being calculated for four out of six consecutive injections (excluding the first and second one); (b) the ODH catalyst was regenerated after every sixth repeated injection runs; (c) the “i” subscript refers to the n-butane, butene, and Coax species under consideration.
Catalysts 14 00505 g004aCatalysts 14 00505 g004b
Figure 5. Comparison between average experimental results and model predictions. Data points for three repeats are reported (standard deviation = ±1.2%).
Figure 5. Comparison between average experimental results and model predictions. Data points for three repeats are reported (standard deviation = ±1.2%).
Catalysts 14 00505 g005
Table 1. Summary of Characterization Results for Supports and Catalysts.
Table 1. Summary of Characterization Results for Supports and Catalysts.
SampleSBET (m2/g)Dpore (Å)H2 Uptake (cm3 STP/g)Reducible V (%)NH3 Uptake (cm3 STP/g)Kdes (mmole/gcat.min)Edes (KJ/mole)
γAl2O3208109--6.031.117 ± 0.7874.43 ± 2.54
5 wt%V/γAl2O31771059.223.028.221.151 ± 0.8969.32 ± 2.12
MgO−γAl2O3 (1:1)152164--4.341.615 ± 0.7148.17 ± 2.14
5 wt% V/MgO−γAl2O31988912.423.635.561.165 ± 0.9676.97 ± 2.02
Table 2. Product Distribution Obtained from Sequential Injections of n-Butane in Oxidative Dehydrogenation Experiments Using the 5 wt%V/MgO-γAl2O3 Catalyst at 10 s and Various Reaction Temperatures.
Table 2. Product Distribution Obtained from Sequential Injections of n-Butane in Oxidative Dehydrogenation Experiments Using the 5 wt%V/MgO-γAl2O3 Catalyst at 10 s and Various Reaction Temperatures.
Temperature (°C)InjectionSelectivity (%)X.C4H10
(%)
Y.C4H8
(%)
COCH4CO2C2H4C2H6C3H6C3H8C4H8
47516.101.638.563.051.3013.552.0263.8022.5614.40
23.571.907.563.451.5313.531.9666.4818.6512.40
32.522.017.213.981.6913.841.8166.9417.7011.85
42.221.976.793.921.7814.571.6567.1117.2211.55
51.981.906.323.841.9415.231.5367.2717.1511.54
61.951.855.663.852.0315.021.4368.2117.0311.61
50016.350.3210.621.300.075.310.6475.3932.9624.85
22.240.417.221.360.075.570.5584.5827.5122.72
31.560.466.071.690.106.010.3885.7326.0621.82
41.420.524.621.910.106.060.3286.8426.1022.20
51.280.514.001.920.105.970.3186.9126.0522.38
61.150.493.551.900.115.760.2987.1425.7622.34
52517.190.3712.031.800.086.880.7270.9233.5123.77
22.750.508.871.670.099.050.6876.3829.0622.20
32.020.597.822.180.129.360.5077.4128.4722.04
41.900.706.202.570.1310.100.4377.9727.6921.59
51.770.715.562.660.1510.500.4378.2227.3521.39
61.630.705.032.700.1610.680.4178.6927.8521.92
55017.610.3314.381.990.0710.340.6664.6237.3624.14
23.280.479.362.530.0811.950.6471.6932.7923.51
32.420.578.742.700.1212.530.4872.4331.1722.58
42.140.667.642.810.1212.980.4073.2430.8422.59
51.930.686.373.030.1413.550.4173.8930.2122.32
61.830.685.313.230.1613.950.4174.4429.6522.07
Table 3. Adsorption Parameters of Different Species. Tm = 515 °C.
Table 3. Adsorption Parameters of Different Species. Tm = 515 °C.
Parameter (atm−1)Estimated Value with 95% Confidence SpansParameter (KJmol−1)Estimated Value with 95% Confidence Spans
K C 4 H 10 0 0.83 ± 0.03 Δ H C 4 H 10 31.2 ± 0.72
K C 4 H 8 0 0.41 ± 0.012 Δ H C 4 H 8 60.8 ± 1.20
K COX 0 0.50 ± 0.018 Δ H COX 54.0 ± 1.05
Table 4. Summary of Intrinsic Kinetic Parameters, with 95% Confidence Intervals (CIs), for the Proposed Kinetic Model when Using the 5 wt% V/MgO−γAl2O3 Catalyst.
Table 4. Summary of Intrinsic Kinetic Parameters, with 95% Confidence Intervals (CIs), for the Proposed Kinetic Model when Using the 5 wt% V/MgO−γAl2O3 Catalyst.
ParametersValue95% CICorrelation Matrix
k 1 * 0 k 2 * 0 k 3 * 0 E1E2E3
k 1 * 0 2.64 × 10−5±2.66 × 10−7 1
k 2 * 0 4.87 × 10−7±4.97 × 10−9 −0.411
k 3 * 0 2.23 × 10−7±4.57 × 10−9 0.40−0.88 1
E189±8.92−0.19−0.25 0.231
E242±4.87−0.36 0.86−0.79−0.28 1
E344±5.34−0.39 0.86−0.89−0.21 0.75 1
m576
DOF570
Table 5. Comparison of Activation Energies for Main BODH Product Formation.
Table 5. Comparison of Activation Energies for Main BODH Product Formation.
CatalystsActivation Energies of Formation (KJ/mole)Reference
C4-OlefinsCarbon Oxides
VOx/MgO−γAl2O38942 a44 bPresent Study
VOx/CeO2−γAl2O390.2105.5 a81 b[30]
V2O5/Al/Mg88.637.4 a45.4 b[31]
VOx/Al2O370.265 a81.3 b[32]
a Formation from N-butane. b Formation from C4-olefins.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bin Sulayman, A.; de Lasa, H. A Kinetic Model for Catalytic N-Butane Oxidative Dehydrogenation under Oxygen-Free Reaction Conditions in a Fluidized CREC Riser Simulator. Catalysts 2024, 14, 505. https://doi.org/10.3390/catal14080505

AMA Style

Bin Sulayman A, de Lasa H. A Kinetic Model for Catalytic N-Butane Oxidative Dehydrogenation under Oxygen-Free Reaction Conditions in a Fluidized CREC Riser Simulator. Catalysts. 2024; 14(8):505. https://doi.org/10.3390/catal14080505

Chicago/Turabian Style

Bin Sulayman, Abdulhamid, and Hugo de Lasa. 2024. "A Kinetic Model for Catalytic N-Butane Oxidative Dehydrogenation under Oxygen-Free Reaction Conditions in a Fluidized CREC Riser Simulator" Catalysts 14, no. 8: 505. https://doi.org/10.3390/catal14080505

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop