Next Article in Journal
Green Fabrication of ZnO Nanoparticles and ZnO/rGO Nanocomposites from Algerian Date Syrup Extract: Synthesis, Characterization, and Augmented Photocatalytic Efficiency in Methylene Blue Degradation
Previous Article in Journal
Platinum Alloys for Methanol Oxidation Electrocatalysis: Reaction Mechanism and Rational Design of Catalysts with Exceptional Activity and Stability
Previous Article in Special Issue
Sodium Methoxide Catalysed One-Pot Glycidol Synthesis via Trans-Esterification between Glycerol and Dimethyl Carbonate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Kinetic Modeling of the Direct Dimethyl Ether (DME) Synthesis over Hybrid Multi-Site Catalysts

1
Unit of Chemical-Physics Fundamentals in Chemical Engineering, Faculty of Science and Technology for Sustainable Development and One Health, University Campus Bio-Medico of Rome, Via Alvaro del Portillo 21, 00128 Rome, Italy
2
CNR-ITAE, Istituto di Tecnologie Avanzate per l’Energia “Nicola Giordano”, Via S. Lucia sopra Contesse 5, 98126 Messina, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(1), 61; https://doi.org/10.3390/catal14010061
Submission received: 20 December 2023 / Revised: 4 January 2024 / Accepted: 11 January 2024 / Published: 13 January 2024

Abstract

:
This paper deals with the proposition of a kinetic model for the direct synthesis of DME via CO2 hydrogenation in view of the necessary optimization of the catalytic system, reactor design, and process strategy. Despite the fact that DME synthesis is typically treated as a mere combination of two separated catalytic steps (i.e., methanol synthesis and methanol dehydration), the model analysis is now proposed by taking into account the improvements related to the process running over a hybrid catalyst in a rational integration of the two catalytic steps, with boundary conditions properly assumed from the thermodynamics of direct DME synthesis. Specifically, the CO2 activation step at the metal–oxide interface in the presence of ZrO2 has been described for the first time through the introduction of an ad hoc mechanism based on solid assumptions from inherent studies in the literature. The kinetic modeling was investigated in a tubular fixed-bed reactor operating from 200 to 260 °C between 1 and 50 bar as a function of a gas hourly space velocity ranging from 2500 to 60,000 NL/kgcat/h, in a stoichiometric CO2/H2 feed mixture of 1:3 v/v. A well-detailed elementary mechanism was used to predict the CO2 conversion rate and identify the key reaction pathways, starting with the analysis of the implicated reactions and corresponding kinetic mechanisms and expressions, and finally estimating the main parameters based on an appropriate modeling of test conditions.

Graphical Abstract

1. Introduction

In consideration of the growing concern for climate change, the search for novel development models based on renewable energy sources with the potential to gradually replace fossil fuels is in the spotlight. On this road to progressive decarbonization, dimethyl ether (DME), the simplest of the ether compounds (chemical formula CH3OCH3), has been recently proposed as a very promising alternative fuel [1,2,3], being characterized by its high cetane number, low auto-ignition temperature, and near-zero emissions of particulate matter, carbon monoxide, and unburned hydrocarbons [4,5]. It also presents chemical and physical properties similar to LPG, enabling the use of existing storage and distribution infrastructure [6]. Additionally, DME can serve as a feedstock for the production of olefins via the DTO (DME-to-olefins) reaction pathway, which is characterized by high yields and conversions [7,8,9].
However, the real appeal of this compound lies in its potential synthesis reaction (Equation (1)), since recent research frontiers aim to employ CO2 and H2 as reactants [10,11]:
2 C O 2 + 6 H 2 C H 3 O C H 3 + 3 H 2 O ( Δ H 0 = 122   k J / m o l )
Using captured CO2 and green H2 produced via electrolysis, it is, therefore, possible to configure DME as an e-fuel directly produced in a sustainable CCUS (carbon capture utilization and storage) technology chain [12,13]. Furthermore, even though this aspect is still under investigation, DME can be steam-reformed, making it a potential circular hydrogen carrier [14]. For this to be feasible, however, it is necessary to make the synthesis process more efficient and cost-effective. Currently, DME is primarily produced from synthesis gas through an indirect method involving the use of two reactors in a typical cascade process [15]. In the first reactor, methanol synthesis is performed on a ternary metal–oxide catalyst (typically CZA, which is CuO-ZnO-Al2O3), while the second reactor allows the dehydration of the previously formed methanol over an acidic catalytic bed (usually γ-Al2O3 or a zeolite sample, such as HZSM-5) [10,16,17,18]. The real challenge is to achieve effective process intensification by implementing direct synthesis in a single reactor, using CO2 as a feedstock, thus allowing the valorization of a thermodynamically “inert” compound, as well as reducing operating and investment costs [19,20]. Certainly, the main concern is the development of a multi-functional catalyst wherein a long-range inter-dispersion among metallic, oxide, and acidic functionalities must be achieved to catalyze both the methanol synthesis and DME production steps [14,21]. Indeed, by hybridizing the catalytic system for direct DME synthesis, the thermodynamic limitations behind the CO2 hydrogenation process stopping the formation of methanol can be overcome, even if unavoidable water generation is a key contributor to possible catalyst deactivation mechanisms, such as copper sintering or migration of weak acidic sites [22,23,24].
Given the potential and critical issues associated with the direct synthesis of dimethyl ether from hydrogen and carbon dioxide, this work aims at investigating the performance, on a laboratory scale, of a previously optimized hybrid CuO-ZnO-ZrO2/HZSM-5 catalytic system [22] by developing a kinetic model suitable to predict its experimental behavior in a wide set of operating conditions. Particularly, the exchange of alumina (a key component in the benchmark catalyst for methanol synthesis) with zirconia to enhance CO2 adsorption, water tolerance, and heat management [25] has been limitedly studied from a microkinetic point of view [26,27]. Therefore, this knowledge gap inspired an ad hoc 18-parameter model based on three cascade steps (i.e., CO2 hydrogenation to methanol, methanol dehydration, and rWGS) occurring in one pot over a hybrid catalytic system. This model aims at explaining a dual-site mechanism on two different adsorption sites by claiming that the CO2 activation step at the metal–oxide interface is crucial prior to undergoing hydrogenation to methanol. Acidic functionality is then necessary for the subsequent dehydration into DME. As a consequence, the corresponding kinetic expressions were derived and appropriately integrated with all the reactions involved in the process. This resulted in a comprehensive model for the direct synthesis of DME, which led to the assessment of new mechanistic clues about the activity–selectivity pattern of the catalyst under investigation as necessary for further optimization of its design.

2. Experimental Section

This section is divided into two parts: synthesis of the hybrid catalyst and validation of its performance through a fixed-bed reactor.

2.1. Catalyst Synthesis

The hybrid catalyst was synthesized through the gel oxalate coprecipitation method, as elsewhere described [28]. Firstly, the nitrate precursors, namely Cu(NO3)2·2.5H2O, Zn(NO3)2·6H2O, and ZrO(NO3)2·nH2O (in a relative Cu/Zn/Zr atomic ratio of 60/30/10), were dissolved in an ethanolic solution and added, at room temperature in a slow and constant mode, to a vigorously stirred ethanolic solution containing oxalic acid (20 wt.% in excess with respect to the stoichiometric amount necessary to precipitate the metal precursors) and the finely dispersed HZSM-5 zeolite (Si/Al = 23 mol/mol) in a weight ratio of 1:1 with respect to the final oxide composition of the catalyst. The precipitate was stirred for 3 h, then aged overnight. The catalyst was filtered, dried at 90 °C for 16 h, and then calcined at 350 °C for 4 h.

2.2. Catalyst Testing

In order to assess the activity of the hybrid catalyst, CO2 hydrogenation to DME was studied in a full kinetic regime, according to conditions designed to rule out the control of diffusional phenomena [11,17]. All the catalytic measurements were carried out in a fixed-bed stainless steel reactor with an internal diameter of 4 mm and a length of 21 mm, loaded with 250 mg of catalyst, and housed within a stainless-steel rod to minimize temperature gradients. The 40–70 mesh (210–420 μm) fraction was selected as a suitable compromise between reactor pressure drop and catalyst particle mechanical resistance. Catalytic tests were carried out at a total pressure of 1, 10, 30, or 50 bar in a temperature range between 200 and 260 °C and gas hourly space velocity of 2500, 8800, and 60,000 NL/kgcat/h, employing a feed mixture of CO2/H2/N2 in a 3:9:1 v/v ratio. Before each test, the catalyst underwent an in situ reduction at 300 °C for 1 h using pure hydrogen flow at atmospheric pressure. The reaction stream was analyzed via a gas chromatograph equipped with a three-column separation system connected to a flame ionization detector (FID) and a thermal conductivity detector (TCD). No formation of methane or higher hydrocarbons was detected under the applied conditions. Quantitative data were expressed in terms of CO2 conversion ( X C O 2 ), DME yield ( Y D M E ) and DME and CO selectivity ( S D M E and S C O ), as defined in Equations (2)–(5):
X C O 2 = F C O 2 i n F C O 2 o u t F C O 2 i n
Y D M E = F D M E o u t F C O 2 i n
S D M E = F D M E o u t F D M E o u t + F M e t h o u t + F C O o u t
S C O = F C O o u t F D M E o u t + F M e t h o u t + F C O o u t
In Equations (2)–(5), F i i n and F i o u t are the inlet and the outlet molar flow rates of component i , respectively. Conversion and selectivity data were determined by both internal standard and mass-balance methods, averaging three independent measurements with an accuracy of ±3%.

3. Kinetic Modeling

The development of a kinetic model for direct CO2-to-DME hydrogenation was carried out in two steps:
  • Analysis of the implicated reactions and corresponding kinetic mechanisms and expressions.
  • Estimation of missing parameters based on appropriate modeling of test conditions.

3.1. Reaction Kinetics Investigation

In direct DME synthesis from CO2 and H2, three reactions (Equations (6)–(8)) are involved [10,29,30,31]:
  • CO2 hydrogenation to methanol (CHM):
    C O 2 + 3 H 2 C H 3 O H + H 2 O ( Δ H 0 = 49   k J / m o l )
  • Methanol dehydration to DME (MDD):
    2 C H 3 O H C H 3 O C H 3 + H 2 O ( Δ H 0 = 23   k J / m o l )
  • Reverse water–gas shift (rWGS):
    C O 2 + H 2 C O + H 2 O ( Δ H 0 = + 41   k J / m o l )
CO hydrogenation to DME is not considered since it can be obtained from the linear combination of Equations (7) and (8), and is also significantly inhibited by the production of formate, an intermediate that strongly binds copper active sites [32,33]. For each reaction, the kinetic mechanism was reconstructed, identifying the rate-determining step (r.d.s.) [34]. As for methanol dehydration and rWGS, previously elaborated mechanisms from the literature were adopted due to the conventional nature of the employed catalyst (HZSM-5 and Cu-based systems, respectively) [35,36], while an ad hoc kinetic mechanism was developed for CO2 hydrogenation, based on experimental evidence [37]. Finally, the kinetic expression of the reaction rate was derived for each mechanism, assuming that all elementary steps, except the r.d.s., were at thermodynamic equilibrium and adopting the Langmuir–Hinshelwood–Hougen–Watson (LHHW) model [38]. In light of the moderate operating conditions in terms of pressure and temperature, the gas phase was assumed to be ideal, with a unitary fugacity coefficient (see Appendix A). A “dual site” model is thus proposed, distinguishing two different active sites on the hybrid catalyst [37,39]: the metal–oxide site ( σ ) and the zeolitic acid site ( σ ~ ) .

3.1.1. CO2 Hydrogenation Mechanism

The introduction of zirconia in the catalyst composition, instead of the alumina present in the benchmark catalyst, has a limited history in the literature related to kinetic modeling [26]. Thus, an ad hoc mechanism was developed to further explore its specific behavior. More in detail, the CO2 activation step is a highly debated aspect, with two generally proposed alternatives [40,41]: the formate route, based on the formation of HCOO intermediate, and the rWGS route, which involves the conversion of CO2 into CO. Although an unambiguous agreement on the exact mechanism does not exist yet, most of the recent experimental findings suggest a leaning towards the formate route, as described by Equations (9a)–(9j):
( 3 x )   H 2 + 2 σ 2 H σ
C O 2 + σ C O 2 σ
C O 2 σ + H σ C H O 2 σ + σ r . d . s .
C H O 2 σ + H σ C H 2 O 2 σ + σ
C H 2 O 2 σ + H σ C H 2 O ( O H ) σ + σ
C H 2 O O H σ + H σ C H 2 O σ + H 2 O σ
H 2 O σ H 2 O + σ
C H 2 O σ + H σ C H 3 O σ + σ
C H 3 O σ + H σ C H 3 O H σ + σ
C H 3 O H σ C H 3 O H + σ
According to this route, one molecule of CO2 activated on a partially oxidized Cu(δ+) site (9b) initially reacts with one atom of hydrogen previously activated on a metallic site (9a), giving rise to a formate species (9c). The latter then reacts with other activated hydrogen atoms in a cascade sequence to produce dioxomethylene species (9d), formaldehyde (9f), methoxy species (9f), and finally methanol (9i) in the last hydrogenation step.

3.1.2. Methanol Dehydration Mechanism

The methanol dehydration reaction strongly relies on the zeolitic component of the catalyst, considering that density, strength, accessibility, and nature of the acid sites (Brønsted or Lewis) are all features capable of influencing the DME productivity [22,37,42]. Several kinetic models have been proposed in the literature, primarily differing from the zeolite topology [26]. However, in terms of reaction mechanism, most studies have converged on identifying two main alternative pathways [42,43]:
Associative pathway, which involves the formation of DME from two adsorbed methanol molecules.
The dpathway is characterized by the adsorption of a methanol molecule followed by its dissociation into a methoxyl group, which then reacts with another methanol molecule to produce DME.
Overall, due to the relatively low operating temperatures (far below 400 °C) and the mesoporous structure of the hybrid catalyst utilized [25], it can be concluded that the predominant pathway is the associative one [44]. Concerning the mechanism over a HZSM-5 zeolite, Lu et al. [35] proposed the following sequence of elementary reactions (Equations (10a)–(10d)):
( 2 x )   C H 3 O H + σ ~ C H 3 O H σ ~
C H 3 O H σ ~ C H 2 σ ~ + H 2 O
C H 2 σ ~ + C H 3 O H σ ~ C H 3 O C H 3 σ ~ + σ ~ r . d . s .
C H 3 O C H 3 σ ~ C H 3 O C H 3 + σ ~

3.1.3. rWGS Mechanism

The rWGS reaction mechanism can be reconstructed based on its counterpart reaction. More specifically, WGS predominantly occurs on copper sites (at the interface with other metal oxides), meaning that the introduction of zirconia can act as a promoter, enhancing the activity of the existing sites without inducing any further reaction [45,46]. It is therefore possible to refer to some mechanisms already found in the literature that identify the carboxyl route as the predominant pathway for the WGS reaction [36,47], as exemplified by the set of Equations (11a)–(11g):
H 2 O + σ H 2 O σ
H 2 O σ + σ O H σ + H σ r . d . s .
C O + σ C O σ
C O σ + O H σ C O O H σ + σ
C O O H σ + σ C O 2 σ + H σ
2 H σ H 2 + 2 σ
C O 2 σ C O 2 + σ

3.2. Dual-Site Adsorption on the Hybrid Catalyst

While the final dehydration step of methanol to DME is known to be mediated by the zeolitic acidic site [15,16,18], an adequate concentration of both metallic and metal–oxide sites is claimed to be somehow necessary for the primary methanol formation [32,33]. In particular, it should be noted that the adsorption of CO2 can even occur on bare oxides, while the dissociation of H2 more efficiently takes place on metallic surfaces; however, a possible activation over metal–oxide surfaces is not ruled out [40,41]. Therefore, this model proposes the use of a unique metal–oxide site for both CO2 and H2 adsorption, assumed to be the active center in CO2 hydrogenation to methanol and in rWGS reaction. Accordingly, the rate equations proposed for this dual-site LHHW mechanism are summarized in Equations (12)–(14):
r C H M = K c i n C H M P C O 2 P H 2 1 / 2 P C H 3 O H P H 2 O K e q C H M P H 2 5 / 2 1 + K C O P C O + K C O 2 P C O 2 + K C H 3 O H P C H 3 O H + K H 2 O P H 2 O + K H 2 P H 2 2
r M D D = K c i n M D D P C H 3 O H 2 P C H 3 O C H 3 P H 2 O K e q M D D 1 + K H 2 O P H 2 O + K C H 3 O H P C H 3 O H + K C H 3 O C H 3 P C H 3 O C H 3 2
r r W G S = K c i n r W G S P H 2 O P C O P C O 2 P H 2 K e q r W G S 1 + K C O P C O + K C O 2 P C O 2 + K C H 3 O H P C H 3 O H + K H 2 O P H 2 O + K H 2 P H 2 2
Remarkably, since both the steps of CO2 hydrogenation to methanol and rWGS are characterized by a competition for the same adsorption site ( σ ), the denominators of Equations (12) and (14) include the terms for CO adsorption ( K C O P C O ) and CH3OH adsorption ( K C H 3 O H P C H 3 O H ), respectively, as inhibition contributions for the corresponding reactions.

3.3. Parameter Estimation Method

In order to validate the kinetic expressions, it is necessary to describe the laboratory setup via a predictive model, which is crucial for parameter estimation. In this light, the fixed-bed reactor was assimilated to an isothermal and isobaric PFR (plug flow reactor). The experimental setup was designed to minimize the impact of internal and external diffusional resistances, which were consequently neglected. Therefore, one-dimensional steady-state mass balances (Equations (15) and (16)) were written for each compound along the axial coordinate:
d G i d z = 1 L j = 1 3 ν i j r j
G i z = 0 = y i i n G T O T i n
In Equation (15), L represents the reactor length, r j is the rate of reaction j (Equations (12)–(14)) and ν i j is the stoichiometric coefficient of compound i in reaction j . In Equation (16), y i i n is the inlet molar fraction of component i and G T O T i n is the total inlet gas hourly space velocity. The balances are, indeed, expressed in terms of space velocity of each compound, noted as G i , which is here defined as the ratio between the molar flow rate F i and the total weight of the catalyst M c (Equation (17)):
G i = F i M c
It is important to note that the kinetic expressions of reaction rates (present in Equation (15)) incorporate temperature-dependent equilibrium, kinetic, and adsorption constants. While the former can be derived from the existing literature, being only subject to the thermodynamics of the reactions involved (see Table 1) [48], the latter two constants are specific to the developed catalyst and require experimental data regression for determination. More precisely, the Arrhenius and Van ‘t Hoff equations express their temperature dependence and were reparametrized (see Equations (18) and (19)) to improve numerical stability in fitting and decrease numerical correlation between estimated parameters [49,50]:
K c i n j = k c i n j exp E A j T r i f T R 1
K i = k i exp Δ H a d s i T r i f T R 1
In Equations (18) and (19), T r i f is a reference temperature, assumed to be equal to the experimental mean value of 503.15 K; T R denotes the reactor temperature; k c i n j and k i are pre-exponential factors; and E A j and Δ H a d s i represent energetic terms. The latter were made dimensionless, according to Equations (20) and (21) (where R is the universal gas constant), to decrease differences in magnitude orders and reduce computational costs [51]:
E A j = E A j T r i f R
Δ H a d s i = Δ H a d s i T r i f R
Both pre-exponential factor and energetic terms were estimated by minimizing the weighted sum ϕ of square relative errors between experimental molar fractions Y i k and model outputs Y ^ i k (Equation (22)) [52]:
ϕ = k = 1 N p i w i Y i k Y ^ i k Y i k 2
In Equation (22), N p represents the number of experimental tests, while w i is the weight assigned to compound i . Optimization was carried out by applying a Matlab algorithm based on the fmincon function to find the minimum value, providing a set of 100 experimental data points as input. The starting point was varied in different trials, looking at achieving a global minimum, while an additional set of 60 data points was used to validate the outcomes obtained.

4. Results and Discussion

4.1. Experimental Results

A sensitivity analysis was carried out to evaluate the influence of temperature and pressure on product distribution, as depicted in Figure 1.
On the whole, the results are consistent with the thermodynamic analysis of the system [10,31,53]. High pressure is preferable during the DME synthesis reaction as it leads to a decrease in the number of moles. On the other hand, an increase in temperature favors CO2 conversion due to its positive kinetic effect on the reaction rate. However, it simultaneously decreases DME selectivity while raising CO selectivity, in agreement with the thermal tonality both of DME synthesis (exo-) and rWGS (endo-). Although a plateau in DME and CO selectivity was observed above 240 °C for runs performed both at 30 and 50 bar, a further increase in reaction temperature unfortunately triggers severe catalyst deactivation. Indeed, a larger amount of water formed above those conditions dramatically determines morphological changes in the catalyst surface, prompting a drastic activity decay (ca. 20% with respect to the initial activity) in a few hours of time on stream (<20 h) [17,22]. Therefore, these findings not only indicate that the process performance is governed by kinetics [15], but also suggest an optimal compromise between CO2 conversion and DME selectivity at moderately high pressures (30 or 50 bar) and intermediate temperatures (<260 °C), so as to limit any issue related to catalyst lifetime.

4.2. Reaction Kinetics Investigation Results

Falling within the typical range of values reported for similar computational models [26], the parameters resulting from the estimation procedure are listed in Table 2, with Figure 2 illustrating the comparison between experimental data and model outputs.
The model provided satisfactory results in calculating the molar fractions of CO2 and H2, with an average relative error of 1.5%, which is comparable to the chromatographic tolerance. Moreover, despite an average relative error of around 40%, the model suitably also works in the prediction of the molar fractions of DME and CO, considering a minimal absolute error in the order of 0.001. Only in the case of the other compounds (i.e., CH3OH or H2O), the large number of estimated parameters (which is equal to 18) cannot determine a lower discrepancy in the model fitting, so that a relatively small error in absolute terms has a strong impact in relative terms.
In any case, even the highest deviation recorded for large methanol concentrations (see Figure 2f) cannot be considered a relevant issue since the simulation model targets the one-pot synthesis of DME and the methanol selectivity always attains relatively low values (below 10%) under all the experimental conditions applied. More importantly, these results evidently confirmed previous mechanistic clues on the catalytic behavior, with positive effects on the process performance played by an increase in space velocity or reaction pressure. However, further optimization of the hybrid catalyst design is required for higher CO2 conversion rates below 220 °C, where CO formation can be minimized. Finally, not only the catalyst formulation but also the reactor configuration should be properly designed in order to tackle the issues related to water formed during the process, dramatically affecting catalyst lifetime and process performance.

5. Conclusions

A comprehensive 18-parameter kinetic model was developed for the direct hydrogenation of CO2 to DME over a hybrid CuO-ZnO-ZrO2/HZSM-5 catalytic system, previously optimized in terms of composition, preparation method, and tuning of physico-chemical properties. The model was validated in a wide set of reaction pressure, temperature, and space velocity, clearly showing a reliable fitting of the estimated parameters, within an average absolute error of 10−4, both for the process reactants (i.e., CO2, H2) and the target product (i.e., DME), with the correspondent experimental values. The accuracy of the model confirmed previous findings on the multi-site nature of the studied process, mainly following the formate route, according to which carbon dioxide is activated at the metal–oxide interface, then reacting with activated hydrogen migrating via spillover from metallic copper sites. Afterwards, formate species are hydrogenated to methanol, which, in line with an associative pathway, quickly turns into DME by condensation with another adsorbed methanol molecule over the acidic sites of the zeolite. Overall, the adequate accuracy in the prediction of the compositions related to the most abundant process compounds makes the model promising for further implementation within a plant simulation, prompting an estimation of the costs associated with the direct DME synthesis process for an assessment of its actual convenience on an industrial scale.

Author Contributions

Conceptualization, L.M., V.P. and F.F.; methodology, A.D., A.B. and L.M.; software, A.D.; validation, A.D., A.B. and L.M.; formal analysis, A.D.; investigation, S.T., M.S. and C.C.; resources, V.P. and G.B.; data curation, A.D.; writing—original draft preparation, A.D.; writing—review and editing, A.D., A.B., L.M. and G.B.; supervision, G.B.; project administration, V.P.; funding acquisition, G.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the European Union’s Horizon Europe research and innovation program under grant agreement No. 101058540 (project PLASTICE).

Data Availability Statement

Data is contained within the article.

Acknowledgments

The authors also acknowledge Francesco Arena from the University of Messina, Dip. Ingegneria, for the fruitful discussion about the boundary conditions of the kinetic model.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A. Fugacity Coefficient Evaluation

When working with real mixtures, it is crucial to determine the fugacity coefficient ν i of each component to measure its deviation from the ideal state. This can be performed by taking advantage of the fact that ln ν i is the partial molar property of ln ν m , where ν m is the fugacity coefficient of the mixture, as described in Equation (A1) [54]:
ln ν i = ln ν m j i y j ln ν m y j P , T , y k j i
In Equation (A1), y i denotes the molar fraction of compound i , while ν m was computed through the two-term virial equation of state (Equations (A2)) [55]:
ln ν m = B m P R T
In Equation (A2), P is the reaction pressure, T is the reaction temperature, and R is the universal gas constant, whereas B m denotes the second virial coefficient of the mixture, which can be evaluated from binary coefficients B i j , as in Equation (A3):
B m = i = 1 7 j = 1 7 y i y j B i j
The binary coefficients B i j can be calculated through Tsonopoulos’s correlation (Equations (A4)–(A7)), which extends the virial equation to polar fluids [56]:
B i j P C i j R T C i j = f 0 T R i j + ω i j f 1 T R i j + f 2 T R i j           w h e r e       T R i j = T T C i j
f 0 T R i j = 0.1445 0.33 T R i j 0.1385 T R i j 2 0.0121 T R i j 3 0.000607 T R i j 8
f 1 T R i j = 0.0637 + 0.331 T R i j 2 0.423 T R i j 3 0.008 T R i j 8
f 2 T R i j = a i j T R i j 6 b i j T R i j 8
The terms f ( 0 ) , f ( 1 ) and f ( 2 ) represent the impact of intermolecular forces from spherically symmetrical molecules, molecular orientation, and polarity, respectively. All the critical binary properties figuring in Equation (A4) are evaluated from the critical properties (i.e., critical volume v C , critical temperature T C , and critical pressure P C ) and acentric factors ω of pure components, as detailed in Equations (A8)–(A11) [57]:
v C i j = v C i 1 3 + v C j 1 3 3 8
T C i j = T C i T C j  
P C i j = T C i j 2 v C i j P C i v C i T C i + P C j v C j T C j
ω i j = ω i + ω j 2
The a i j and b i j parameters in Equation (A7) measure polar interactions and are exclusively non-zero for pairs of polar fluids. In such instances, they are assumed to be equal to the average of the constants of pure components, as outlined in Equations (A12) and (A13) [55]:
a i j = a i + a j 2
b i j = b i + b j 2
In the specific case analyzed, the only polar fluids considered are water, methanol, and dimethyl ether, whose parameters are listed in Table A1.
Table A1. a and b parameters for Equations (A12) and (A13) [55].
Table A1. a and b parameters for Equations (A12) and (A13) [55].
Compound
a
b
Water0.02790.0229
Methanol0.08780.0560
Dimethyl Ether−0.015130
Based on the above, in Table A2 are reported the fugacity coefficients as determined under the most severe operating conditions (i.e., 200 °C and 50 bar). The analysis indicates that, in nearly all cases, these coefficients are very close to one, resulting in a negligible correction to the partial pressure. Moreover, although nitrogen shows a measurable deviation from its ideal behavior, its fugacity is not present in the kinetic expressions of the reaction rates (Equations (12)–(14)). Therefore, it was preferred to maintain the validity of the ideal gas hypothesis against a relatively modest enhancement in the model accuracy only achievable behind an increased computational complexity.
Table A2. Fugacity coefficients for different compositions, evaluated at 200 °C and 50 bar.
Table A2. Fugacity coefficients for different compositions, evaluated at 200 °C and 50 bar.
y C O 2 y H 2 y D M E y C O y M e t h y H 2 O y N 2 ϕ C O 2 ϕ H 2 ϕ D M E ϕ C O ϕ M e t h ϕ H 2 O ϕ N 2
0.23 0.69 ---- 0.08 0.97 1.03 ---- 0.24
0.21 0.67 - 0.02 - 0.02 0.08 0.97 1.03 - 1.02 -0.96 0.24
0.21 0.67 0.005 0.007 0.002 0.016 0.09 0.97 1.03 0.98 1.01 0.95 0.96 0.24
0.20 0.63 0.01 0.03 0.01 0.03 0.09 0.97 1.03 0.97 1.01 0.94 0.95 0.24
0.19 0.62 0.02 0.02 0.01 0.06 0.08 0.96 1.04 0.97 1.02 0.93 0.94 0.24

References

  1. Fleisch, T.H.; Basu, A.; Gradassi, M.J.; Masin, J.G. Dimethyl Ether: A Fuel for the 21st Century. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 1997; pp. 117–125. [Google Scholar]
  2. Matzen, M.; Demirel, Y. Methanol and Dimethyl Ether from Renewable Hydrogen and Carbon Dioxide: Alternative Fuels Production and Life-Cycle Assessment. J. Clean. Prod. 2016, 139, 1068–1077. [Google Scholar] [CrossRef]
  3. Makoś, P.; Słupek, E.; Sobczak, J.; Zabrocki, D.; Hupka, J.; Rogala, A. Dimethyl Ether (DME) as Potential Environmental Friendly Fuel. E3S Web Conf. 2019, 116, 00048. [Google Scholar] [CrossRef]
  4. Park, S.H.; Lee, C.S. Applicability of Dimethyl Ether (DME) in a Compression Ignition Engine as an Alternative Fuel. Energy Convers. Manag. 2014, 86, 848–863. [Google Scholar] [CrossRef]
  5. Park, S.H.; Lee, C.S. Combustion Performance and Emission Reduction Characteristics of Automotive DME Engine System. Prog. Energy Combust. Sci. 2013, 39, 147–168. [Google Scholar] [CrossRef]
  6. Semelsberger, T.A.; Borup, R.L.; Greene, H.L. Dimethyl Ether (DME) as an Alternative Fuel. J. Power Sources 2006, 156, 497–511. [Google Scholar] [CrossRef]
  7. Khadzhiev, S.N.; Kolesnichenko, N.V.; Ezhova, N.N. Manufacturing of Lower Olefins from Natural Gas through Methanol and Its Derivatives (Review). Pet. Chem. 2008, 48, 325–334. [Google Scholar] [CrossRef]
  8. Cordero-Lanzac, T.; Martínez, C.; Aguayo, A.T.; Castaño, P.; Bilbao, J.; Corma, A. Activation of N-Pentane While Prolonging HZSM-5 Catalyst Lifetime during Its Combined Reaction with Methanol or Dimethyl Ether. Catal. Today 2022, 383, 320–329. [Google Scholar] [CrossRef]
  9. Pérez-Uriarte, P.; Ateka, A.; Aguayo, A.T.; Gayubo, A.G.; Bilbao, J. Kinetic Model for the Reaction of DME to Olefins over a HZSM-5 Zeolite Catalyst. Chem. Eng. J. 2016, 302, 801–810. [Google Scholar] [CrossRef]
  10. Catizzone, E.; Bonura, G.; Migliori, M.; Frusteri, F.; Giordano, G. CO2 Recycling to Dimethyl Ether: State-of-the-Art and Perspectives. Molecules 2017, 23, 31–58. [Google Scholar] [CrossRef]
  11. Bonura, G.; Cordaro, M.; Spadaro, L.; Cannilla, C.; Arena, F.; Frusteri, F. Hybrid Cu–ZnO–ZrO2/H-ZSM5 System for the Direct Synthesis of DME by CO2 Hydrogenation. Appl. Catal. B 2013, 140–141, 16–24. [Google Scholar] [CrossRef]
  12. Wawrzyńczak, D.; Majchrzak-Kuceba, I.; Pevida, C.; Bonura, G.; Nogueira, R.; De Falco, M. The Carbon Chain in Carbon Dioxide Industrial Utilization Technologies: A Case Study; Taylor & Francis: Abingdon, UK, 2023. [Google Scholar]
  13. Sankhe, S.; Krishna, S.V.M.; Juturu, R.M.; Subrahmanyam, C. Power-to-X (PtX) Technologies and Their Potential Role in the Transition towards a Fossil-Free Energy Future: A Review of EFuels Synthesis and Direct Air Capture (DAC) Technology. In Proceedings of the International Conference on Automotive Materials and Manufacturing AMM 2023, Pune, India, 1–3 June 2023; SAE Technical Paper 2023-28-1333. pp. 1–5. [Google Scholar] [CrossRef]
  14. Catizzone, E.; Freda, C.; Braccio, G.; Frusteri, F.; Bonura, G. Dimethyl Ether as Circular Hydrogen Carrier: Catalytic Aspects of Hydrogenation/Dehydrogenation Steps. J. Energy Chem. 2021, 58, 55–77. [Google Scholar] [CrossRef]
  15. Azizi, Z.; Rezaeimanesh, M.; Tohidian, T.; Rahimpour, M.R. Dimethyl Ether: A Review of Technologies and Production Challenges. Chem. Eng. Process. Process Intensif. 2014, 82, 150–172. [Google Scholar] [CrossRef]
  16. Mondal, U.; Yadav, G.D. Perspective of Dimethyl Ether as Fuel: Part I. Catalysis. J. CO2 Util. 2019, 32, 299–320. [Google Scholar] [CrossRef]
  17. Bonura, G.; Migliori, M.; Frusteri, L.; Cannilla, C.; Catizzone, E.; Giordano, G.; Frusteri, F. Acidity Control of Zeolite Functionality on Activity and Stability of Hybrid Catalysts during DME Production via CO2 Hydrogenation. J. CO2 Util. 2018, 24, 398–406. [Google Scholar] [CrossRef]
  18. Tokay, K.C.; Dogu, T.; Dogu, G. Dimethyl Ether Synthesis over Alumina Based Catalysts. Chem. Eng. J. 2012, 184, 278–285. [Google Scholar] [CrossRef]
  19. Akhoondi, A.; Osman, A.I.; Alizadeh Eslami, A. Direct Catalytic Production of Dimethyl Ether from CO and CO2: A Review. Synth. Sinter. 2021, 1, 105–125. [Google Scholar] [CrossRef]
  20. Chen, W.H.; Hsu, C.L.; Wang, X.D. Thermodynamic Approach and Comparison of Two-Step and Single Step DME (Dimethyl Ether) Syntheses with Carbon Dioxide Utilization. Energy 2016, 109, 326–340. [Google Scholar] [CrossRef]
  21. Banivaheb, S.; Pitter, S.; Delgado, K.H.; Rubin, M.; Sauer, J.; Dittmeyer, R. Recent Progress in Direct DME Synthesis and Potential of Bifunctional Catalysts. Chem. Ing. Tech. 2022, 94, 240–255. [Google Scholar] [CrossRef]
  22. Bonura, G.; Cannilla, C.; Frusteri, L.; Catizzone, E.; Todaro, S.; Migliori, M.; Giordano, G.; Frusteri, F. Interaction Effects between CuO-ZnO-ZrO2 Methanol Phase and Zeolite Surface Affecting Stability of Hybrid Systems during One-Step CO2 Hydrogenation to DME. Catal. Today 2020, 345, 175–182. [Google Scholar] [CrossRef]
  23. Ren, S.; Shoemaker, W.R.; Wang, X.; Shang, Z.; Klinghoffer, N.; Li, S.; Yu, M.; He, X.; White, T.A.; Liang, X. Highly Active and Selective Cu-ZnO Based Catalyst for Methanol and Dimethyl Ether Synthesis via CO2 Hydrogenation. Fuel 2019, 239, 1125–1133. [Google Scholar] [CrossRef]
  24. Abu-Dahrieh, J.; Rooney, D.; Goguet, A.; Saih, Y. Activity and Deactivation Studies for Direct Dimethyl Ether Synthesis Using CuO–ZnO–Al2O3 with NH4ZSM-5, HZSM-5 or γ-Al2O3. Chem. Eng. J. 2012, 203, 201–211. [Google Scholar] [CrossRef]
  25. Frusteri, L.; Bonura, G.; Cannilla, C.; Todaro, S.; Giordano, G.; Migliori, M.; Frusteri, F. Promoting Direct CO2 Conversion to DME over Zeolite-Based Hybrid Catalysts. Pet. Chem. 2020, 60, 508–515. [Google Scholar] [CrossRef]
  26. Ateka, A.; Rodriguez-Vega, P.; Ereña, J.; Aguayo, A.T.; Bilbao, J. Kinetic Modeling and Reactor Design of the Direct Synthesis of Dimethyl Ether for CO2 Valorization. A Review. Fuel 2022, 327, 125148. [Google Scholar] [CrossRef]
  27. Hadipour, A.; Sohrabi, M. Synthesis of some bifunctional catalysts and determination of kinetic parameters for direct conversion of syngas to dimethyl ether. Chem. Eng. J. 2008, 137, 294–301. [Google Scholar] [CrossRef]
  28. Frusteri, F.; Cordaro, M.; Cannilla, C.; Bonura, G. Multifunctionality of Cu–ZnO–ZrO2/H-ZSM5 Catalysts for the One-Step CO2-to-DME Hydrogenation Reaction. Appl. Catal. B 2015, 162, 57–65. [Google Scholar] [CrossRef]
  29. Delgado Otalvaro, N.; Bilir, P.G.; Herrera Delgado, K.; Pitter, S.; Sauer, J. Kinetics of the Direct DME Synthesis: State of the Art and Comprehensive Comparison of Semi-Mechanistic, Data-Based and Hybrid Modeling Approaches. Catalysts 2022, 12, 347. [Google Scholar] [CrossRef]
  30. Naik, S.P.; Ryu, T.; Bui, V.; Miller, J.D.; Drinnan, N.B.; Zmierczak, W. Synthesis of DME from CO2/H2 Gas Mixture. Chem. Eng. J. 2011, 167, 362–368. [Google Scholar] [CrossRef]
  31. De Falco, M.; Capocelli, M.; Centi, G. Dimethyl Ether Production from CO2 Rich Feedstocks in a One-Step Process: Thermodynamic Evaluation and Reactor Simulation. Chem. Eng. J. 2016, 294, 400–409. [Google Scholar] [CrossRef]
  32. Studt, F.; Behrens, M.; Kunkes, E.L.; Thomas, N.; Zander, S.; Tarasov, A.; Schumann, J.; Frei, E.; Varley, J.B.; Abild-Pedersen, F.; et al. The Mechanism of CO and CO2 Hydrogenation to Methanol over Cu-Based Catalysts. ChemCatChem 2015, 7, 1105–1111. [Google Scholar] [CrossRef]
  33. Cao, A.; Wang, Z.; Li, H.; Elnabawy, A.O.; Nørskov, J.K. New Insights on CO and CO2 Hydrogenation for Methanol Synthesis: The Key Role of Adsorbate-Adsorbate Interactions on Cu and the Highly Active MgO-Cu Interface. J. Catal. 2021, 400, 325–331. [Google Scholar] [CrossRef]
  34. Campbell, C.T. Finding the Rate-Determining Step in a Mechanism: Comparing DeDonder Relations with the “Degree of Rate Control”. J. Catal. 2001, 204, 520–524. [Google Scholar] [CrossRef]
  35. Lu, W.Z.; Teng, L.H.; Xiao, W. De Simulation and Experiment Study of Dimethyl Ether Synthesis from Syngas in a Fluidized-Bed Reactor. Chem. Eng. Sci. 2004, 59, 5455–5464. [Google Scholar] [CrossRef]
  36. Li, Z.; Li, N.; Wang, N.; Zhou, B.; Yin, P.; Song, B.; Yu, J.; Yang, Y. Mechanism Investigations on Water Gas Shift Reaction over Cu(111), Cu(100), and Cu(211) Surfaces. ACS Omega 2022, 7, 3514–3521. [Google Scholar] [CrossRef] [PubMed]
  37. Bonura, G.; Todaro, S.; Frusteri, L.; Majchrzak-Kucęba, I.; Wawrzyńczak, D.; Pászti, Z.; Tálas, E.; Tompos, A.; Ferenc, L.; Solt, H.; et al. Inside the Reaction Mechanism of Direct CO2 Conversion to DME over Zeolite-Based Hybrid Catalysts. Appl. Catal. B 2021, 294, 120255. [Google Scholar] [CrossRef]
  38. Motagamwala, A.H.; Dumesic, J.A. Microkinetic Modeling: A Tool for Rational Catalyst Design. Chem. Rev. 2021, 121, 1049–1076. [Google Scholar] [CrossRef]
  39. Saravanan, K.; Ham, H.; Tsubaki, N.; Bae, J.W. Recent progress for direct synthesis of dimethyl ether from syngas on the heterogeneous bifunctional hybrid catalysts. Appl. Catal. B Environ. 2017, 217, 494–522. [Google Scholar] [CrossRef]
  40. Din, I.U.; Shaharun, M.S.; Alotaibi, M.A.; Alharthi, A.I.; Naeem, A. Recent Developments on Heterogeneous Catalytic CO2 Reduction to Methanol. J. CO2 Util. 2019, 34, 20–33. [Google Scholar] [CrossRef]
  41. Guil-López, R.; Mota, N.; Llorente, J.; Millán, E.; Pawelec, B.; Fierro, J.L.G.; Navarro, R.M. Methanol Synthesis from CO2: A Review of the Latest Developments in Heterogeneous Catalysis. Materials 2019, 12, 3902. [Google Scholar] [CrossRef]
  42. Blaszkowski, S.R.; van Santen, R.A. The Mechanism of Dimethyl Ether Formation from Methanol Catalyzed by Zeolitic Protons. J. Am. Chem. Soc. 1996, 118, 5152–5153. [Google Scholar] [CrossRef]
  43. Jones, A.J.; Iglesia, E. Kinetic, Spectroscopic, and Theoretical Assessment of Associative and Dissociative Methanol Dehydration Routes in Zeolites. Angew. Chem. 2014, 126, 12373–12377. [Google Scholar] [CrossRef]
  44. Ghorbanpour, A.; Rimer, J.D.; Grabow, L.C. Computational Assessment of the Dominant Factors Governing the Mechanism of Methanol Dehydration over H-ZSM-5 with Heterogeneous Aluminum Distribution. ACS Catal. 2016, 6, 2287–2298. [Google Scholar] [CrossRef]
  45. Tang, Q.-L.; Liu, Z.-P. Identification of the Active Cu Phase in the Water−Gas Shift Reaction over Cu/ZrO2 from First Principles. J. Phys. Chem. C 2010, 114, 8423–8430. [Google Scholar] [CrossRef]
  46. Madon, R.J.; Braden, D.; Kandoi, S.; Nagel, P.; Mavrikakis, M.; Dumesic, J.A. Microkinetic Analysis and Mechanism of the Water Gas Shift Reaction over Copper Catalysts. J. Catal. 2011, 281, 1–11. [Google Scholar] [CrossRef]
  47. Lacerda de Oliveira Campos, B.; Herrera Delgado, K.; Wild, S.; Studt, F.; Pitter, S.; Sauer, J. Surface Reaction Kinetics of the Methanol Synthesis and the Water Gas Shift Reaction on Cu/ZnO/Al2O3. React. Chem. Eng. 2021, 6, 868–887. [Google Scholar] [CrossRef]
  48. Behloul, C.R.; Commenge, J.-M.; Castel, C. Simulation of Reactors under Different Thermal Regimes and Study of the Internal Diffusional Limitation in a Fixed-Bed Reactor for the Direct Synthesis of Dimethyl Ether from a CO2-Rich Input Mixture and H2. Ind. Eng. Chem. Res. 2021, 60, 1602–1623. [Google Scholar] [CrossRef]
  49. Agarwal, A.K.; Brisk, M.L. Sequential Experimental Design for Precise Parameter Estimation. 1. Use of Reparameterization. Ind. Eng. Chem. Process Des. Dev. 1985, 24, 203–207. [Google Scholar] [CrossRef]
  50. Pritchard, D.J.; Bacon, D.W. Statistical Assessment of Chemical Kinetic Models. Chem. Eng. Sci. 1975, 30, 567–574. [Google Scholar] [CrossRef]
  51. Espie, D.M.; Macchietto, S. Nonlinear Transformations for Parameter Estimation. Ind. Eng. Chem. Res. 1988, 27, 2175–2179. [Google Scholar] [CrossRef]
  52. Ereña, J.; Sierra, I.; Aguayo, A.T.; Ateka, A.; Olazar, M.; Bilbao, J. Kinetic Modelling of Dimethyl Ether Synthesis from (H2 + CO2) by Considering Catalyst Deactivation. Chem. Eng. J. 2011, 174, 660–667. [Google Scholar] [CrossRef]
  53. Ahmad, K.; Upadhyayula, S. Greenhouse Gas CO2 Hydrogenation to Fuels: A Thermodynamic Analysis. Environ. Prog. Sustain. Energy 2019, 38, 98–111. [Google Scholar] [CrossRef]
  54. Marrelli, L. Termodinamica Degli Equilibri Di Fasi Fluide; Edizioni Efesto: Rome, Italy, 2017. [Google Scholar]
  55. De Santis, R. Introduzione al Calcolo Degli Equilibri Di Fasi Fluide; ESA: Paris, France, 1991. [Google Scholar]
  56. Tsonopoulos, C. An Empirical Correlation of Second Virial Coefficients. AIChE J. 1974, 20, 263–272. [Google Scholar] [CrossRef]
  57. Chueh, P.L.; Prausnitz, J.M. Vapor-Liquid Equilibria at High Pressures. Vapor-Phase Fugacity Coefficients in Nonpolar and Quantum-Gas Mixtures. Ind. Eng. Chem. Fundam. 1967, 6, 492–498. [Google Scholar] [CrossRef]
Figure 1. Influence of temperature and pressure on CO2 conversion (a), DME yield (b), DME selectivity (c), and CO selectivity (d), evaluated with an inlet flow rate of 8800 NL/(kgcat h).
Figure 1. Influence of temperature and pressure on CO2 conversion (a), DME yield (b), DME selectivity (c), and CO selectivity (d), evaluated with an inlet flow rate of 8800 NL/(kgcat h).
Catalysts 14 00061 g001
Figure 2. Comparison of experimental data (along the x-axis) and model results (along the y-axis) for the mole fractions of CO2 (a), H2 (b), DME (c), CO (d), H2O (e), and methanol (f).
Figure 2. Comparison of experimental data (along the x-axis) and model results (along the y-axis) for the mole fractions of CO2 (a), H2 (b), DME (c), CO (d), H2O (e), and methanol (f).
Catalysts 14 00061 g002
Table 1. Expressions of equilibrium constants’ dependence on temperature [48].
Table 1. Expressions of equilibrium constants’ dependence on temperature [48].
Parameter Temperature Dependence
K e q C H M 10 3066 T 10.592         [ b a r 2 ]
K e q M D D exp 4019 T + 3.707 log T 2.783 10 3 T + + 3.8 10 7 T 2 6.561 10 4 T 3 26.64
K e q r W G S 10 2073 T + 2.029
Table 2. Estimated parameters for the proposed kinetic model and corresponding confidence interval with a level of significance equal to 0.05.
Table 2. Estimated parameters for the proposed kinetic model and corresponding confidence interval with a level of significance equal to 0.05.
Parameter Unit Pre-Exponential Factor
( k c i n j   or   k i )
Energetic Term
( E A j   or Δ H a d s i )
K c i n C H M [ m o l g c a t 1 s 1 b a r 3 / 2 ] ( 2.56 ± 0.05 ) 10 4   44.95 ± 1.75
K c i n M D D [ m o l g c a t 1 s 1 b a r 2 ] 0.26 ± 0.01 99.73 ± 1.21
K c i n r W G S [ m o l g c a t 1 s 1 b a r 2 ] ( 6.16 ± 0.12 ) 10 3   39.52 ± 1.71
K H 2 [ b a r 1 ] ( 8.54 ± 0.32 ) 10 2 33.57 ± 1.38
K C O 2 [ b a r 1 ] 9.56 ± 0.10   14.23 ± 0.93
K C O [ b a r 1 ] 8.97 ± 0.49 10.41 ± 0.40
K H 2 O [ b a r 1 ] 8.02 ± 0.43 22.95 ± 0.86
K C H 3 O H [ b a r 1 ] 5.70 ± 0.29 7.07 ± 0.11
K C H 3 O C H 3 [ b a r 1 ] 5.20 ± 0.10 60.53 ± 0.66
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

D’Ambrosio, A.; Bertino, A.; Todaro, S.; Santoro, M.; Cannilla, C.; Frusteri, F.; Bonura, G.; Mazzeo, L.; Piemonte, V. Kinetic Modeling of the Direct Dimethyl Ether (DME) Synthesis over Hybrid Multi-Site Catalysts. Catalysts 2024, 14, 61. https://doi.org/10.3390/catal14010061

AMA Style

D’Ambrosio A, Bertino A, Todaro S, Santoro M, Cannilla C, Frusteri F, Bonura G, Mazzeo L, Piemonte V. Kinetic Modeling of the Direct Dimethyl Ether (DME) Synthesis over Hybrid Multi-Site Catalysts. Catalysts. 2024; 14(1):61. https://doi.org/10.3390/catal14010061

Chicago/Turabian Style

D’Ambrosio, Antonio, Alice Bertino, Serena Todaro, Mariarita Santoro, Catia Cannilla, Francesco Frusteri, Giuseppe Bonura, Leone Mazzeo, and Vincenzo Piemonte. 2024. "Kinetic Modeling of the Direct Dimethyl Ether (DME) Synthesis over Hybrid Multi-Site Catalysts" Catalysts 14, no. 1: 61. https://doi.org/10.3390/catal14010061

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