Masoud Safari,
Ali Haghtalab* and
Farzaneh Arabpour Roghabadi
Faculty of Chemical Engineering, Department of Process, Tarbiat Modares University, P.O. Box: 14115-143, Tehran, Iran. E-mail: haghtala@modares.ac.ir
First published on 1st August 2023
To enhance the catalyst performance of Fischer–Tropsch synthesis (FTS), removing the mass-transfer restriction in the catalysis synthesis is essential. Although the core–shell nanostructures can improve the activity and stability of the catalyst, they can restrict the reactants' diffusion towards the active sites and the transfer of the products from these sites in FTS. Creating an adequate porosity between the core and the outer shell of the catalyst structure can tackle this issue. In this work, the synthesized cobalt-based nano-catalyst is encapsulated with two shells and a middle porous shell. The first shell is a carbon shell at the core of the catalyst derived from ZIF-67, the second one is the outer shell of ceria, and the middle porous shell is formed by removing the sacrificial silica shell through the etching technique. The characterization and performance tests represent significant evidence of the etching treatment's impact on the FTS catalyst performance. Besides, molecular dynamics simulation is also utilized to clarify its effect. The FTS catalytic performance is enhanced more than 2 times with the etched catalyst versus the catalyst without it at 17.5 bar and a (H2/CO) ratio of 1.2. In addition, not only does the etched catalyst with high porosity play the role of a nanoreactor and intensify its catalytic performance, but it also has higher stability.
There is a wide range of researches on FTS cobalt-based catalysts that utilize the core–shell structure. A prominent achievement is represented by encapsulating the cobalt nanoparticles with two shells consisting of an inert nanolayer of carbon which encapsulates and immobilizes cobalt nanoparticles in the core and the main outer shell is formed by inorganic oxides like SiO2 or TiO2.1,4,9–12 The carbon nanolayers can be derived from a metal–organic framework (MOF) such as ZIF-67, which can provide the highest cobalt dispersion due to the confinement effect of the carbon framework.1,9 The inorganic oxides have presented significant features like the simplicity of fabrication, tunability of formation, cost-effectiveness, tunable morphology, adjustable sizes, extensive surface area, and pore volume. So, these metal oxides can be nominated as the best material for synthesizing the outer shell.8
Although the inorganic oxide shell has many suitable attributes, its multi-layer matrix structure reduces the reactants' and products' mass transfer rate through the FTS process. Besides, in core–shell structure, the outer shell may cover some of the active Co sites, leading to a decline in the accessibility of the reactants, and then a diminution occurs in the CO conversion.9
In this work, a novel catalyst structure is introduced to tackle this issue. First of all, according to recent studies,1,9,10 ZIF-67 is synthesized and the nano-porous carbon layers are derived from it through the thermal steps. Nano-porous carbon layers encapsulate the active cobalt sites and provide a suitable dispersion at the catalyst's core. The core is encapsulated by two shells that are SiO2 and CeO2, where the middle one is employed as a sacrificial shell and removed at the end step of catalyst preparation, providing adequate hollow space to facilitate the diffusion of reactants towards the active sites or transfer of products out of them. The outer shell, CeO2, is formed to involve the mentioned features of the inorganic oxide. Following the initial evaluation of the characteristics of the synthesized catalysts, we studied their performance in the FTS. Accordingly, for a better understanding of the porous shell impacts on the performance of the catalyst in the FTS, the diffusion of the reactants and products is evaluated by employing molecular dynamic (MD) simulation.
(1) |
During the H2-TPD test, 0.07 g catalyst is perfectly reduced at 475 °C for 12 h by utilizing a H2 flow, then cooled to 100 °C in the H2 stream and held at this temperature in the Ar ambient for one hour. Later, the temperature is raised to 475 °C with 5 °C min−1 to desorb all the hydrogen. Eventually, the sample is re-oxidized at 475 °C by pulses of O2 in He as the carrier gas. By measuring the amount of oxygen consumption during the oxygen titration and based on the percentage of reduction, the DOR of each sample and the corrected dispersion can be evaluated, respectively.14,15 The test is applied in a U-tube quartz reactor with the quantachrome CHEMBET-3000 unit, which incorporates a thermal conductivity detector (TCD).
N2 adsorption–desorption isotherms and Brunauer–Emmett–Teller surface area are obtained using Micromeritics Tristar 3020 on degassed samples. TEM imaging is implemented on Talos F200X (FEI) microscopes operated at 200 kV. The cobalt crystalline diameter (dTEM) can be calculated from the TEM results by using eqn (3) if at least 200 nanoparticles are considered.
(2) |
To implement the field emission scanning electron microscope (FE-SEM), the TESCAN MIRA3 is utilized.
The obtained results from the MD simulations depend on the choice of the interatomic force field. In this study, the interatomic force field for various compounds is based on universal force field (UFF) potential.21 The computation phase is implemented by determining the force field for molecules in the simulation box. Then the atomic diffusion ratio/time is computed. These coefficients averaged per 10000-time steps to eliminate numerical fluctuations. Because of the computational point, the mean square displacement (MSD) parameter should be calculated first to estimate the diffusion coefficient. It is the most common measure of the spatial extent of random motion and can be thought of as measuring the portion of the system estimated by the random walker.
In the second simulation section, after the equilibrium phase occurs in simulated structures, their thermal behavior is estimated. In this step of the calculations, the Green–Kubo approach is used to describe the thermal behavior of simulated samples.22,23 Technically, thermal calculation in the current study is performed with a micro-canonical (NVE) ensemble as the following equation:
(3) |
In this equation kB, V, and S define the Boltzmann constant (1.380649 × 10−23 J K−1), volume, and heat vector, respectively. At each simulation time step in LAMMPS computational package, the heat current of atoms can be estimated by using velocity and stress tensor. The mean of the heat current autocorrelation function is represented by <S(t)·S(0)>.
Table 1 represents MD simulation settings in current computational research. Structurally, various atomic membranes are modeled inside MD simulation boxes. x (out shell thickness), y (porous shell/SiO2 shell diameter), and z (core diameter) constants in these membranes are the principal structural parameters which are listed in Table 2. In this model, the potential energy for an arbitrary geometry of a molecule is expressed as a superposition of bonded (Eval) and non-bonded (Enb) interactions as eqn (4):24
E = Eval + Enb | (4) |
MD Simulation parameter | Value/settings |
---|---|
Box length | 500 × 500 × 500 nm3 |
Initial temperature | 300 K |
Time step | 0.001 fs |
Used ensembles | NVT |
Temperature damping ratio | 0.1 |
Simulation time | 10 ns |
Membrane ID | x (nm) | y (nm) | z (nm) |
---|---|---|---|
NC–V catalyst | 8 | 360 | 330 |
NC catalyst | 8 | 355 | 330 |
The bonded interactions consist of bond strength, bond-angle bend, dihedral angle torsion, and inversion terms. In this model, non-bond interaction between atoms is modeled by the Lennard–Jones (LJ) potential as below formalism:25
(5) |
The variables qi and qj denote the partial charge of ionic species of i and j atoms, respectively, and ε0 is the permittivity of the vacuum. The first variable of the van der Waals term prevents the overlap of the atoms and molecules, and the second term of this potential represents the short-range repulsive interaction and dipole-induced attractive interaction, respectively. In this equation, rij is the distance between two atoms, rc is the cutoff radius, and εij and σij are the energy and length scales. Both εij and σij parameters depend on the type of atoms in the simulated structure. The length scale parameter and energy for various atoms of defined compounds are listed in Table 3.21 Finally, simple harmonic oscillator equations describe the bond-based stretch in the united atom model.
Atoms | ε (kcal mol−1) | σ (angstrom) |
---|---|---|
C | 0.105 | 3.851 |
Ce | 0.013 | 3.556 |
Co | 0.014 | 2.82 |
H | 0.044 | 2.886 |
O | 0.060 | 3.500 |
For computations of the atom's motion through time, Newton's second law at the atomic level is used as the gradient of the potential (force field) function in the following equation,24
(6) |
The Association of defined formulations is done by the Velocity-Verlet method to integrate Newton's law.26,27
Then, after the gas and liquid separation by two traps, they can be analyzed by utilizing an Agilent 7890A refinery gas analyzer (RGA) and an Agilent 7890A detailed hydrocarbon analyzer (DHA), respectively. By applying the results of the analyzers, the CO conversion (XCO) can be evaluated, and then by subtracting the fraction of CO that is used for the formation of C1 to C4 products (hydrocarbons with one to four carbons), the C5+ (hydrocarbons with more than five carbons) selectivity can be calculated. In addition, the turnover frequency (TOF) can be evaluated by dividing the mole of converted CO by the mole of active sites per reaction time.28
(7) |
(8) |
SC5+(%) = 100 − (SC1 + SC2 + SC3 + SC4) | (9) |
Fig. 2 (a) XRD patterns, (b) H2-TPR profiles, (c) TGA profiles of the NC–V (blue line) and NC (red line) catalysts. |
Sample | SBET (m2 g−1) | VP (cm3 g−1) | PD (nm) | Da (nm) | Db (nm) | DOR (%) | Dispersion (%) |
---|---|---|---|---|---|---|---|
NC–V catalyst | 414 | 0.445 | 4.3 | 6.76 | 8.84 | 78.1 | 11.24 |
NC catalyst | 219 | 0.118 | 2.16 | 7 | 9.87 | 65.73 | 10.07 |
Then, the temperature-programmed reduction in H2 (H2-TPR) is conducted to study the reducibility of the metallic species of the catalyst. Fig. 2b illustrates the H2-TPR curves of both catalysts, which depict three prominent peaks. Following Jacobs et al.29 the first peak indicates the reduction of Co3O4 to CoO, and the second one expresses the reduction of metallic Co. The NC–V catalyst's H2-TPR patterns demonstrate that these reduction peaks are approximately placed at low temperatures of 343 and 413 °C, respectively. Furthermore, a tiny peak can be detected for this catalyst at a lower temperature of 300 °C. As a result, silica removal paves the way for reducing the Co nanoparticles (NPs) with lower energy consumption. In contrast, the peaks of the H2-TPR of the NC catalyst initiate at 373 °C and follow by the second one at 438 °C. In addition, due to the existence of cobalt silicate which is hardly reduced, the third peak in the NC catalyst occurs at a higher temperature of 785 °C with a broader band compared with the NC–V catalyst which is considered another evidence of the silica removal during the etching step of the NC–V catalyst synthesis and shows the silica removal's role on facilitating the catalyst reduction at lower temperatures. As Fig. 2b shows, the third peak for the NC–V catalyst is at 629 °C. Furthermore, following the Trovarelli30 report, the majority of ceria reduction happens at the temperature range of 600–850 °C, so the third peak also covers the ceria reduction while it reveals that the reduction of ceria in the NC catalyst in the presence of silica is harder than the etched catalyst (NC–V). Besides, the low TPR peaks of the NC–V catalyst imply the facilitated reduction owing to the etching treatment.
In addition, to reveal the impact of the etching treatment on the catalyst performance, the degree of reduction (DOR) and the rate of Co active sites dispersion is evaluated by the hydrogen temperature programmed desorption (H2-TPD) and oxygen titration. The results of these parameters are rendered in Table 4 which represents that the NC–V catalyst has higher DOR and appropriate dispersion of Co active sites compared to the NC catalyst. Thus, these results indicate that the silica coating at the NC catalyst engages more Co active sites than the NC–V catalyst. Besides, the average size of cobalt particles is also derived from the H2-TPD test.14 In both characterizations (XRD, and H2-TPD), there is not a significant distinction between the size of both catalysts, indicating no destructive impact of the etching step on the particle size of the catalyst.
Moreover, the thermal gravimetric analysis (TGA) is implemented to study the range of temperature decomposition of both catalysts, representing a stunning distinction between them as shown in Fig. 2c. The NC–V catalyst loses its weight by less than 5% until 400 °C while NC catalyst loses more than 11%. Besides, the NC catalyst is decomposed roughly to 40% of its initial mass until 1000 °C; on the contrary, the NC–V catalyst does not lose more weight than 6% during the same raising temperature condition which is in good agreement with the other works.31,32 The weight loss of the NC catalyst is roughly the same as that of silica-coated catalysts in the other literature.9,33 The TGA results reveal the impact of the internal void space as an insulator on increasing the NC–V catalyst thermal stability. In this regard, to indicate the role of NC–V catalyst's internal void space in its thermal stability, the thermal resistance of both catalysts is compared by computing their thermal conductivity with MD. If the thermal conductivity of the catalyst is lower, the heat transfer to the core of the catalyst will be restricted and the catalyst decomposition will be decreased. The catalyst decomposition occurs when metal oxides' bulk oxygen ions transfer to the surface sites34,35 or the carbon layer oxidizes at a high temperature. Hence, low thermal conductivity can provide suitable resistance versus high-temperature increases.
To compare the textural properties of the two catalyst samples N2-adsorption/desorption is utilized. Both of the catalysts display the type IV adsorption isotherm with type H2 hysteresis that locates at P/P0 ≈ 0.45 and P/P0 ≈ 0.54 for the NC catalyst and NC–V catalyst, respectively (Fig. 3a and b). Furthermore, Table 4 represents the other textural properties of catalysts, the Brunauer–Emmett–Teller area (SBET), and the total pore volume (VP) of the NC–V catalyst is placed at a higher level than the NC catalyst. Also, the pore diameter (PD) of the NC–V catalyst with 4.3 nm is higher than the NC catalyst with 2.16 nm. These results reveal chemical treatment's impact on modifying the catalyst's porosity. Besides, diagrams of the pore volume versus the pore diameter in Fig. 4a and b exhibits that roughly 70% of NC–V catalyst particles are mesoporous; on the contrary, 69% of NC catalyst particles are microporous. The microporous particles ratio to the mesoporous particles in the NC–V and NC catalysts is 0.24 and 2.27, respectively. Hence, the NC–V catalyst's mesoporous structure implies the internal porous shell in the etched catalyst.
Fig. 4 BJH adsorption cumulative pore volume (pore size distribution): (a) NC–V catalyst and (b) NC catalyst. |
TEM and FE-SEM/EDX are applied to provide more information on the textural properties of both types of catalysts. Representative TEM images of the catalyst samples are displayed in Fig. 5a and b which reveal internal void spaces in the etched catalyst by red remarking in contrast with the not etched catalyst. Besides, the crystallite size distribution diagram of each catalyst is shown in Fig. 6a and b. In addition, the TEM images of spent catalysts after TOS of 192 h are shown in Fig. 7a and b. After the etching step, the NC–V catalyst images have much brighter areas around the darker central core of Co3O4@C. These very bright areas are related to the removal of a part of the silica layer in the core–shell structure, which is more reflective than the rest of the parts. The etching treatment does not completely remove the silica layer but only randomly removes the silica in this layer, and this area is brighter than the rest of the areas in the TEM image. The phenomenon of cavity formation has also been confirmed by BET analysis, that the surface area of the sample has increased due to the etching effect, which is the reason for the formation of internal void spaces in the NC–V catalyst.
The FE-SEM map data (Fig. S1 and S2 in ESI†) indicates a uniform elemental distribution all over both catalyst samples and gives prominent evidence of silica shell elimination in the NC–V catalysts. Besides, they reveal that the carbon nanolayer is maintained, despite the calcination of the catalyst in the air following Wang et al.10 Besides, the results of imaging methods reconfirm that not only the chemical treatment step does not leave any damage but also plays the role of a catalyst decorator to reveal the catalyst's active sites. Furthermore, the results of the chemical composition of the catalyst samples are expressed in Table 5, which the FE-SEM/EDX and ICP-OES measures. The comparison of Si percentage in both catalysts that are presented in Table 5 highlights the Si elimination through the etching step, although a mild percentage of Si has remained. Due to the removal of the silica coating, more percentage of active cobalt sites can be achieved, and as a result, the etched catalyst (NC–V) represents higher cobalt loading values than the NC catalyst, which is demonstrated in Table 5. The cobalt amounts in the NC catalyst that the ICP-OES reports are roughly 12% lower than the NC–V catalyst. The etching step modifies the catalyst and makes more accessways towards active cobalt sites by removing the majority of the silica coating shell. As Table 5 demonstrates, not only the cobalt NPs are unwrapped during the etching step, but also the ceria particles are mostly flourished.
Sample | Elemental content (FE-SEM/EDX) % | Elemental content (ICP-OES) w% | ||||
---|---|---|---|---|---|---|
Co | Si | Ce | Co | Si | Ce | |
NC–V catalyst | 18.19 | 12.26 | 69.55 | 27.8 | 11.6 | 60.6 |
NC catalyst | 11.5 | 46.17 | 42.33 | 13.7 | 49 | 37.3 |
The results of the MD computations are depicted in Fig. 8a and b. Not only does the diffusion of the reactants in the NC–V catalyst overtake the NC catalyst, but also the diffusion of the hydrocarbon products outside of the NC–V catalyst is higher than in the NC catalyst. In addition, the MD results illustrate a descending pattern for the diffusion values of the hydrocarbon chains in both catalysts, which have occurred by increasing the number of carbon atoms. The descending pattern becomes more noticeable from methane to pentane which then continues with a gradual trend that not only does it bring a simple confirmation of MD simulation correctness by obvious evidence but also compares the catalysts' confinement for the diffusion of the components with different molecular weight.
As Table 6 illustrates, the calculated thermal conductivity of the NC catalyst, 15.24 W m−1 K−1, is roughly 18.5% more than the NC–V catalyst. Besides, the computed heat flux value in the NC catalyst, 0.831 W m2, is higher than the NC–V catalyst with a value of 0.696 W m2. Hence, the MD computations reconfirm that the thermal resistance of the NC–V catalyst is higher than the NC catalyst.
Sample | Heat flux (W m2) | Thermal conductivity (W m−1 K−1) |
---|---|---|
NC–V catalyst | 0.696 | 12.86 |
NC catalyst | 0.831 | 15.24 |
Sample | Pres. (bar) | H2/CO | CO (conv.%) | CTY (10−4 mol CO g−1 Co s−1) | TOF (10−3 s−1) | SC1 (%) | SC2–4 (%) | SCO2 (%) | SC5+ (%) |
---|---|---|---|---|---|---|---|---|---|
NC–V catalyst | 17.5 | 1.2 | 48 | 5.2 | 8.1 | 4.4 | 2.13 | 0.025 | 93.47 |
NC–V catalyst | 20 | 2 | 81 | 12 | 18.5 | 5 | 2.2 | 0.006 | 92.8 |
NC–V catalyst | 25 | 1.6 | 86 | 13 | 19.6 | 4 | 1.58 | 0.01 | 94.42 |
NC catalyst | 17.5 | 1.2 | 18.2 | 2.2 | 3.9 | 7.6 | 7.2 | 0.035 | 85.2 |
NC catalyst | 20 | 2 | 39 | 6.4 | 11.2 | 11 | 6 | 0.01 | 83 |
NC catalyst | 25 | 1.6 | 63 | 10 | 18.1 | 8.8 | 4 | 0.02 | 87.2 |
CO2 is mainly formed by the water gas shift reaction. According to the other studies,36,37 a high tendency to reverse water gas shift reaction for the Co/CeO2 catalysts at the low temperature is observed, which is derived from the low tendency of Co NPs to WGSR and ceria's oxygen vacancies with a high tendency to CO2 adsorption. Hence, during the catalytic tests, a negligible CO2 selectivity is detected, which is lower in the NC–V catalyst than in the NC catalyst (Table 7).
On the other hand, Fig. 9a and b depict, respectively, the CO conversion and SC5+ values for each catalyst versus time on stream (TOS) of 192 h at the two different operating conditions (A: 245 °C, 17.5 bar, H2/CO = 1.2, and GHSV = 800 h−1 – B: 245 °C, 25 bar, H2/CO = 1.6, and GHSV = 800 h−1). It presents that the NC–V catalyst has considerably higher stability in both operating conditions with a noticeable distinction in CO conversion. The NC–V catalyst loses its conversion by approximately 6% which its trend is shown in Fig. 9 by two lines, each of them having a slope of 0.0155 and 0.0207 at (A) and (B) conditions, respectively. In contrast, Fig. 9a exhibits that the NC catalyst performance reduces by 32% and 16% at the (A) and (A) conditions, respectively, with the line's slope of 0.047 and 0.0649. Besides, Fig. 9b indicates that the SC5+ decreases more noticeably in the NC catalyst, with line's slope of 0.0168 and 0.0246 than in the NC–V catalyst, with line's slope of 0.0126 and 0.0123, at the (A) and (B) conditions, respectively.
Etching at the end of catalyst fabrication steps provides an adequate void space between the core and outer shell of CeO2. This nanoreactor framework intensifies the operative pressure and temperature by creating an effective confinement void around the active sites. Besides, not only does the porous shell of this nanoreactor framework supply adequate void space for conducting the reaction effectively, but also the accessibility to the active cobalt sites is significantly facilitated owing to removing the majority cover of the silica shell that coated the active cobalt sites or bonded with them as cobalt silicate. Therefore, through this structure, the entire metal sites at the core might be accessible, and the NC–V catalyst structure's ability to perform FTS is higher in comparison with the core–shell configuration of the NC catalyst. In particular, in comparison with the NC catalyst, at low pressure and with the minimum ratio value of H2/CO, a noticeable CO conversion with high selectivity to C5+ can be achieved in the FTS process in the presence of the NC–V catalyst (Table 7). On the other hand, it is essential to supply high pressure which means high energy consumption to obtain an appropriate CO conversion by the NC catalyst. Thus, even though high pressure leads to higher CO conversion in the NC catalyst, the selectivity grows towards unsuitable products like methane, and a tiny growth occurs for C5+. Furthermore, compared with the NC catalyst, the NC–V catalyst illustrates higher conversion at a higher pressure, and the selectivity to lighter hydrocarbons has a slight drop.
From another point of view, after the FTS process reaches a steady state condition, the products are formed and occupy the pores of the catalyst, which can lead to the reduction of the reactants (H2, CO) accessibility to the active sites and as a result, the CO conversion may decrease. Because the accumulation of the primarily generated hydrocarbons, particularly in the core–shell structure of the NC catalyst, does not allow the reactants to reach the active cobalt sites, the CO conversion is reduced. Therefore, the diffusion value of products has a crucial role in enhancing the catalyst performance.
It should be mentioned that trapping the products in the pores of the catalyst may increase the residence time and lead to chain growth through re-adsorption or secondary reaction, particularly of the smaller chains. Eventually, the high selectivity to the longer hydrocarbon chains could be achieved. Still, in agreement with the literature that applied the encapsulated cobalt-base catalyst by silica,1,9,11 this core–shell structure would restrict the reaching of the reactants (H2, CO) to the active sites. Therefore, although the longer residence time may enhance the longer chain products, it shouldn't be as much to reduce the reaction yield by restricting the conversion that leads to low heavier hydrocarbon formation. The experimental results confirm that the desirable situation can be achieved through the etched catalyst (NC–V). Following the review study of Zhou Wei et al.44 the mesoporosity of the catalyst suppresses the light hydrocarbons. The low selectivity of the NC–V catalyst towards the light hydrocarbons which have two to four carbons (SC2–4%) indicates that the void space can pave the way to enhance the selectivity toward hydrocarbons with more than five carbons (SC5+%) by facilitating the diffusion of reactants and products. Moreover, the higher DOR of the etched catalyst in comparison to the silica-coated catalyst provides a higher percentage of cobalt active sites which leads to higher CHx formation and intensifies the C–C bonds formation to the carbon chain growth significantly. Therefore, lower selectivity to the light hydrocarbons (C2–C4) occurs in the NC–V catalyst versus the NC catalyst.
On the other hand, the stability of the etched catalyst is higher than the without etched catalyst. TGA test indicates a low significant weight loss under high temperatures for the etched catalyst, which reveals its high thermal stability. Besides, the MD computations highlight the only structural distinction of these catalysts, the internal porous shell, which reduces the catalyst's thermal conductivity value at the NC–V catalyst. So, the internal porous shell plays the role of thermal resistance in the etched catalyst which reduces the heat transfer to the core of the catalyst due to the lower thermal conductivity and protects the catalyst from decomposition. Besides, the evaluation of CO conversion variation during TOS of 192 h indicates that the NC–V catalyst has remarkable stability during the FTS process. The TEM image of the spent NC catalyst indicates darker spots than the NC–V catalyst, which reveals the agglomeration of catalyst NPs. In contrast, the TEM image of the NC–V catalyst's spent sample illustrates the same duplicate as the fresh NC–V catalyst, which verifies its high stability against deactivation. As a result, the internal porous shell raises the catalyst's resistance against harsh thermal conditions and improves its stability significantly.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ra04884e |
This journal is © The Royal Society of Chemistry 2023 |