Next Article in Journal
Tailored Ni-MgO Catalysts: Unveiling Temperature-Driven Synergy in CH4-CO2 Reforming
Previous Article in Journal
Ru-Loaded Biphasic TiO2 Nanosheet-Tubes Enriched with Ti3+ Defects and Directionally Deficient Electrons as Highly Efficient Catalysts in Benzene Selective Hydrogenation
Previous Article in Special Issue
Efficient Hydrogen Production from the Aqueous-Phase Reforming of Biomass-Derived Oxygenated Hydrocarbons over an Ultrafine Pt Nanocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessment of Reaction Kinetics for the Dehydrogenation of Perhydro-Dibenzyltoluene Using Mg- and Zn-Modified Pt/Al2O3 Catalysts

by
Rudaviro Garidzirai
,
Phillimon Modisha
* and
Dmitri Bessarabov
*
HySA Infrastructure Centre of Competence, Faculty of Engineering, North-West University, Potchefstroom Campus, Private Bag X6001, Potchefstroom 2520, South Africa
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(1), 32; https://doi.org/10.3390/catal14010032
Submission received: 13 September 2023 / Revised: 10 December 2023 / Accepted: 20 December 2023 / Published: 29 December 2023
(This article belongs to the Special Issue Recent Advances in Heterogeneous Catalysis for Low-Carbon Fuels)

Abstract

:
The catalysts utilized for the dehydrogenation of dibenzyltoluene-based liquid organic hydrogen carriers (LOHCs) remain crucial. The state-of-the-art catalyst for dehydrogenation of dibenzyltoluene-based LOHC still suffers from deactivation and by-product formation. This is crucial in terms of the efficiency of the industrial dehydrogenation plant for hydrogen production, cyclability as well as the cost of replacing the catalyst. The development of catalysts with optimum performance, minimum deactivation and low by-product formation is required to attain the full benefits of the LOHC technology. Therefore, in this study, the effect of Mg and Zn modification on Pt/Al2O3 catalyst is investigated for the catalytic dehydrogenation of perhydro-dibenzyltoluene (H18-DBT). In addition, an assessment of reaction kinetics is also conducted. High dehydrogenation performance was obtained for Mg-doped Pt/Al2O3 using a batch reactor at 300 °C and 6 h reaction time. In this case, the degree of dehydrogenation (dod), productivity and conversion obtained are 100%, 1.84 gH2/gPt/min and 99.9%, respectively. Moreover, the Mg-doped catalyst has resulted in a high turnover frequency (TOF) of 586 min−1 compared to the Zn-doped catalyst (269 min−1) and the undoped catalyst (202 min−1) at the reaction temperature of 300 °C. The amount of by-products increased with an increase in the catalytic activity, with the Pt/Mg-Al2O3 catalyst possessing the highest amount of by-products. The dehydrogenation of H18-DBT followed first-order reaction kinetics. In addition, the activation energy obtained using the Arrhenius model is 102, 130 and 151 kJ/mol for Pt/Al2O3, Pt/Zn-Al2O3 and Pt/Mg-Al2O3, respectively. Although the Mg-doped Pt/Al2O3 shows high activation energy, the higher performance of the catalyst suggests that mass transfer limitations have no major effect on the dehydrogenation reaction under the conditions used.

Graphical Abstract

1. Introduction

Hydrogen is one of the most pursued alternative energy carriers due to its high gravimetric energy density (33 kWh/kg), which is higher than most conventional fuels [1]. It is also a clean energy carrier and a feedstock that has the potential to decarbonize various fossil-based sectors, such as petrochemicals, chemicals, glass, cement, steel, etc. Specifically, Hydrogen has a low density (0.0813 g/L at STP), resulting in low energy per volume; hence, efficient storage methods are needed [2]. Several hydrogen storage technologies, such as high-pressure gas cylinders, liquefied hydrogen, ammonia, metal hydrides and liquid organic hydrogen carriers (LOHCs), have been explored to mitigate the storage challenges [3]. Amongst the different storage mediums, LOHC technology offers numerous advantages over conventional hydrogen storage technologies. This is because hydrogen can be stored on a large scale for long periods of time without losses when using LOHC technology, whilst hydrogen storage in cylinders or as liquified hydrogen has potential losses. Dibenzyltoluene (H0-DBT) has been identified as one of the most promising LOHC molecules for hydrogen storage because of its attractive hydrogen storage properties [4]. Hydrogen is stored by the catalytic hydrogenation of H0-DBT to produce the hydrogen-rich molecule known as perhydrodibenzyltoluene (H18-DBT). Since this is a reversible process, hydrogen is released by catalytic dehydrogenation of H18-DBT to produce dibenzyltoluene as a hydrogen lean molecule. The H0-DBT/H18-DBT LOHC system is non-flammable, not classified as dangerous goods, remains liquid at sub-zero temperatures and has a high storage capacity (6.2 wt %, 57 kg-H2/m3-H18-DBT) and boiling point (380 °C) [4,5]. Other attractive properties of this system include its compatibility with existing infrastructure for fuel, industrial-scale production and cost-effectiveness (3–4 €/kg) [4,5]. Platinum has been the most widely used catalyst for the dehydrogenation of H18-DBT in the production of hydrogen [6,7,8,9,10,11]. However, Pt-based catalysts (e.g., Pt/Al2O3) are prone to deactivation during the dehydrogenation process, which suggests a loss in the activity of the active sites [12,13]. Over time, this will lead to a reduction in the efficiency of the dehydrogenation system as well as the cost implications associated with frequent catalyst replacement. Various efforts have been employed to improve performance and reduce by-products. This includes, but is not limited to, surface pre-treatment of catalysts [6], support modifications [13,14,15] and utilization of different methods of catalyst synthesis [16].
The EU-supported SHERLOHCK consortium has set criteria which dehydrogenation catalysts should meet [17]. This includes a productivity of >3 gH2/gcatalyst/min and low energy demand (<6 kWh/kg-H2) while maintaining high conversion (>90%) and selectivity of >99.8%. The initial catalyst performance screening tests for dehydrogenation of H18-DBT were first conducted by Brückner et al. [3]. The high catalytic activity was reported for Pt/C with a hydrogen yield of 71% after 3.5 h at 270 °C. This was followed by Pt/Al2O3, Pd/C, Pt/SiO2 and Pd/Al2O3, with a degree of dehydrogenation (dod) of 51%, 16%, 10% and 8%, respectively. It can be deduced from the screening results that SiO2 support and Pd-based catalysts are not suitable for H18-DBT dehydrogenation. The work of Modisha et al. [12] also confirmed that Pt has high catalytic performance for dehydrogenation of H18-DBT compared to Pd and Pt–Pd supported on Al2O3. To date, several hydrogenation and dehydrogenation cycles have been attempted and the presence of by-products has been observed in each cycle. The liquid-based by-products (e.g., benzyltoluene and benzylmethylfluorene) are also LOHCs that could subsequently affect the properties (e.g., viscosity, boiling point, etc.) of the parent molecules [18]. Gaseous by-products in the hydrogen effluent, such as CH4, CO and CO2, have been found to be due to the presence of impurities (water and oxygenates) in the LOHC [12,19,20].
The catalyst can be modified to improve its activity and stability and to reduce the formation of by-products. Recently, Shi et al. [6] modified the surface hydroxyl groups and oxygen vacancies of Al2O3 using H2 and O2 dielectric barrier discharge plasma. Both treatments led to decreased Pt size and improved Pt dispersion for Pt/Al2O3 (3 wt %). The hydrogen-treated catalyst showed improved catalytic performance because the modification increased surface hydroxyl groups. There are a few publications that report on the effect of modifiers/dopants on the dehydrogenation of H18-DBT. Auer et al. [14] investigated the effect of sulphur dopant on the dehydrogenation of H18-DBT using 0.3 wt % Pt/Al2O3. The appropriate sulphur loading (0.25 wt %) reduced the formation of by-products and improved the dod from 78% to >90%. Likewise, Chen et al. [15] modified 3 wt % Pt/Al2O3 and Pt/TiO2 using 0.25 wt % sulphur. Doping with sulphur improved the activity of both catalysts; however, Pt/Al2O3 produced more by-products than Pt/TiO2. Furthermore, a sulphur loading of 0.5 wt % on Pt/Al2O3 showed higher dod (95%) compared to undoped Pt/Al2O3 with a dod of 81%. Apart from sulphur, other dopants have been considered. For example, Seidel [21] studied dopants such as lithium, potassium and caesium in different ratios for modification of the Pt/Al2O3 catalyst. A dopant to Pt ratio of 5.7 was found to be the optimum, with the lithium dopant showing high dod (85%).
Although the abovementioned strategies have been beneficial, there is still a need to comprehend the reaction kinetics for the dehydrogenation of H18-DBT. Further catalyst development work will be beneficial for the optimization of reaction conditions and, ultimately, the improvement of the reactor design. It is reported that the dehydrogenation of H18-DBT depends on the mass transport of the molecules, which involves the diffusion of reactants and products to and from the catalyst pores [10]. The kinetic models that have been developed for the dehydrogenation of H18-DBT are based on typical Pt/Al2O3 catalysts, but the effect of dopants on the kinetic parameters has not been fully explored. Peters et al. [22] developed a kinetic model for H18-DBT dehydrogenation using 5 wt % Pt/Al2O3 in a fixed bed reactor at temperatures from 290 to 350 °C. First-order reaction kinetics was obtained, with activation energy (Ea) and turnover frequency (TOF) of 117 kJ/mol and 120 s−1, respectively. The Pt/Al2O3 catalyst was also used by Park et al. [23] for the dehydrogenation of H18-DBT at temperatures ranging from 250 to 320 °C using a continuous flow reactor. The reaction rate orders obtained were between 2.3 and 2.4, with an activation energy of 171 kJ/mol. Modisha et al. [18] obtained an activation energy of 201 kJ/mol using a batch reactor. A microchannel reactor indicated improved reaction kinetics compared with conventional reactors [8]. According to the literature, metal doping could lead to changes in the electronic, structural and chemical properties of the catalyst, which affects the catalyst’s performance [14,15,24,25]. In this study, the effects of Mg and Zn dopants on the Pt/Al2O3 catalyst performance and reaction kinetics were investigated for the dehydrogenation of H18-DBT using a batch reactor.

2. Results and Discussion

This section addresses the results of the effect of reaction temperature and dopants on the catalytic dehydrogenation performance of H18-DBT and the dehydrogenation kinetics.

2.1. Evaluation of Catalytic Performance for Dehydrogenation of H18-DBT Using a Batch Reactor

The effect of the dehydrogenation reaction temperature on catalytic performance was evaluated using a batch reactor at temperatures in the range of 260–300 °C. The catalysts were subjected to four dehydrogenation cycles (runs) to determine the catalytic performance. The four dehydrogenation runs (run 1, run 2, run 3 and run 4) were also used to determine the stability of the catalysts. To determine the reaction kinetics for the dehydrogenation of H18-DBT, run 4 was used as it showed stability. Stability in this instance means the activity of catalysts was constant for at least two consecutive runs.
Figure 1 displays the effect of reaction temperature on hydrogen flow, productivity and the dod between 260 and 300 °C for the Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3 catalysts. When the dehydrogenation reaction temperature was increased, there was an increase in hydrogen flow, productivity and dod. This is expected because an increase in temperature increases the rate of reaction, resulting in higher activity. A linear increase was observed for hydrogen flow and productivity in all runs. The activity of the catalysts decreased after the first run, yet some stability was observed between runs 3 and 4 (see also hydrogen flow and dod vs. time in Figure S1). This decrease is due to catalyst deactivation. In run 1, there was no apparent difference in hydrogen flow and productivity between 260 and 280 °C; however, at 290 and 300 °C, slight differences were observed. In runs 2–4 and at temperature ranges of 260–290 °C, the activity of all catalysts was similar, but the Mg-doped catalyst exhibited the highest activity at 300 °C. The fresh Pt/Mg-Al2O3 catalyst produced a 100% dod at 290 °C for run 1, while, at the same temperature, the Pt/Al2O3 and Pt/Zn-Al2O3 catalysts produced a dod of 78%. Pt/Zn-Al2O3 and Pt/Al2O3 did not provide full dehydrogenation at the studied temperatures. After the first run, Pt/Mg-Al2O3 still exhibited the highest dod of 73% compared with values of 60% and 50% observed for Pt/Zn-Al2O3 and Pt/Al2O3, respectively.
The unmodified Pt/Al2O3 used in this work showed a productivity of 1.62 gH2/gPt/min, whereas literature work [14,15] reported a range of 0.6–1.2 gH2/gPt/min. The Mg- and Zn-doped catalysts showed productivities of 1.84 and 1.68 gH2/gPt/min, respectively. This is comparable with the values obtained by Auer et al. [14] and Chen et al. [15]. These authors found that doping the 0.3 wt % Pt/Al2O3 with sulphur resulted in productivities of 2 and 1.5 gH2/gPt/min, respectively. An improved productivity of 2 gH2/gPt/min was obtained with low Pt loading (0.3 wt %) [14] compared to the productivity of 1.84 gH2/gPt/min obtained in this study using 0.5 wt % Pt/Al2O3 (see Figure 2). This is because high Pt dispersion is achieved with low Pt loading, which results in improved catalytic activity [14,15]. Nonetheless, it is considered that the performances recorded in this study are comparable with literature work.
In this study, Mg appears to be the best dopant for the dehydrogenation of H18-DBT. The improved performance of the Pt/Mg-Al2O3 catalyst could be due to the reduction in deactivation by the Mg dopant, as will be expanded upon in the discussion below. The comparison of the productivity of the catalysts prepared in this work and the literature work is presented in Figure 2.
The effects of temperature and dopants on the conversion, selectivity and by-product formation were investigated by analyzing the LOHC samples using gas chromatography single quadrupole mass spectroscopy (GC-SQ-MS). During the dehydrogenation of H18-DBT, the intermediates H12-DBT and H6-DBT are formed and then also become dehydrogenated to produce H0-DBT. Therefore, the dehydrogenation of H18-DBT follows the following order: H18-DBT-3H2 → H12-DBT-3H2 →H6-DBT-3H2 → H0-DBT. Figure 3 shows the total ion chromatogram of the LOHC reaction mixture.
In the dehydrogenation of H18-DBT at 300 °C, fresh and used catalysts (from run 1 to run 4) showed conversions of >90% for Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3 used in this study (Figure 4) (Figure 5). However, the H18-DBT conversion of the used catalyst was significantly low (~40% decrease) at temperatures from 260 °C to 270 °C. This phenomenon can be explained by the fact that the dehydrogenation of H18-DBT is endothermic and sensitive to temperature changes. Although the difference in conversion for all catalysts at the same temperature is less pronounced, the H0-DBT selectivity for the Pt/Mg-Al2O3 catalysts is remarkable. At the reaction temperature of 290 °C, the Pt/Mg-Al2O3, Pt/Al2O3 and Pt/Zn-Al2O3 catalysts showed H0-DBT selectivities of 87%, 72% and 77%, respectively. A comparable selectivity was obtained for all catalysts with an apparent difference at 300 °C. In run 4, Pt/Al2O3 showed the lowest selectivity (35%), while Pt/Zn-Al2O3 and Pt/Mg-Al2O3 showed selectivities of 49% and 62%, respectively, at 300 °C.
The amount of by-products increased with an increase in both the dehydrogenation temperature and the catalyst activity. By-products, in this case, are a combination of high and low boiling point compounds, as reported in our previous work [18]. Table 1 shows that the highest amount of by-products (12 mol %) was produced by the Pt/Mg-Al2O3 catalyst, while Pt/Zn-Al2O3 and Pt/Al2O3 produced only 5 and 1.2 mol %, respectively, at 300 °C. In run 4, the amounts decreased to 0.68%, 2.1% and 1.9 mol % for Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3, respectively. Therefore, a decrease in the amount of by-products can be correlated with low catalytic activity observed in run 4. These by-products are due to the cracking and cyclization of dibenzyltoluene molecules and it was found that the by-products are chemically similar throughout the runs (with the only difference being the quantity). The higher the catalytic reaction, the higher the dod and the formation of aromatic molecules. These aromatic molecules are susceptible to the acidic sites of the catalyst, which provides the cracking function. A combination of side reactions leading to the formation of by-products as well as the intermediates certainly affect the selectivity to H0-DBT.
The formation of by-products increases with an increase in the degree of dehydrogenation over time, as shown in Figure 6. This indicates that as the dehydrogenation reaction takes place, the amount of partially and fully dehydrogenated compounds are prone to side reactions; hence, by-products are formed. In this study, the aliphatic by-products have not been obtained. Figure 7 shows possible reaction mechanisms for H0-DBT C–C bond breaking, which forms low boiling point by-products, and dehydrocyclization of H0-DBT, which forms high boiling point by-products.
The deactivation parameter, X = (Xi − Xf)/Xi × 100, was calculated based on the H18-DBT conversion to determine the percentage deactivation based on the reaction time of 6 h (four runs of 90 min per run). Figure 8 shows the percentage deactivation of the catalysts at 300 °C. The undoped Pt/Al2O3 had the highest deactivation of 7.4%, while Zn-doped and Mg-doped catalysts had deactivations of 4.9% and 1.9%, respectively. This suggests that Mg doping suppresses deactivation in a batch reactor, hence leading to an improved performance.
A summary of the performance of catalysts in a batch reactor is presented in Table 2. The highest activity for all catalysts was observed at 300 °C. At this temperature, the addition of Mg and Zn dopants significantly increased activity, conversion, selectivity and by-products. For example, the dod improved from 82 to 88% for Zn-doped Pt/Al2O3 and from 82 to 100% for Mg-doped Pt/A2O3. Moreover, full dehydrogenation was achieved within 75 min when using the Mg-doped catalysts, whereas undoped and Zn-doped catalysts did not result in full dehydrogenation within 90 min. For all catalysts, a decrease in activity was observed after the first run. The loss of activity can be attributed to catalyst deactivation. The better performance of the Pt/Mg-Al2O3 catalysts can be explained by the low deactivation parameter of 1.9%, which suggests that Mg reduced deactivation. Characterization results obtained from our previous work [13] indicate that Mg reduced the acidity of alumina by 37% and this could have led to better performance than in the cases of Zn-doped and undoped catalysts with similar acidity values (see Table 2). The strong acidic sites inhibit the desorption of products; hence, the ability of Mg to decrease these sites led to improved selectivity to H0-DBT and the ultimate enhancement of performance. Due to the high catalytic activity of Pt/Mg-Al2O3, a high amount of by-products were observed. Therefore, in this study, Pt/Mg-Al2O3 showed improved catalytic performance for the dehydrogenation of H18-DBT compared to Pt/Al2O3 and Pt/Zn-Al2O3. This study is an extension of our previously published work and the physicochemical properties of the catalysts studied in this work can be obtained from our previous work [13].

2.2. Reaction Kinetics for Dehydrogenation of H18-DBT

The effects of temperature on the kinetic parameters (e.g., reaction rate, rate constant, TOF) were then investigated. Experiments were conducted at dehydrogenation temperatures in the range of 260–290 °C and the dod was <70%. This was to avoid back-reactions that occur at high conversions [20].
As depicted in Figure 9, the concentration of H18-DBT decreased with time for all catalysts and the decrease in H18-DBT concentration is a function of temperature. Hence, the concentration gradient becomes steeper with an increase in temperature. At 300 °C, a H18-DBT concentration of 0.84 M was observed for Pt/Mg-Al2O3, while Pt/Zn-Al2O3 and Pt/Al2O3 generated final concentrations of 1.30 and 1.42 M, respectively.
From Equation (5), the rate constants (k), initial rate of reactions (r) and turnover frequencies (TOF) can be calculated using the Equations (1) and (2), as reported by Sotoodeh [28].
TOF is defined as the number of reactant molecules converted by each Pt molecule per unit time.
r = k × C 0
TOF = k × Mwt Pt × C 0 M Pt × D
Here, r is the initial rate of reaction (M min−1 gcat−1), k is the rate constant (min−1), C 0 is the initial concentration of H18-DBT (M), MwtPt is the molecular mass of Pt (g/mol), MPt is the Pt loading (wt %) and D is the Pt dispersion (%).
The dehydrogenation kinetics of H18-DBT followed first-order reaction kinetics, as seen by the linear fit of ln (C/C0) vs. time in Figure 10. This is in accordance with the literature [12,18]. Rate constants were obtained from the slopes of the respective curves and then used to calculate the rates of reaction and TOFs, as shown in Table 3. The activity increases with an increase in temperature. This is due to the collision of molecules induced by high temperatures, which then increases the reaction rate. Undoped Pt/Al2O3 produced higher rate constants at low temperatures (260 and 270 °C) than Pt/Mg-Al2O3 and Pt/Zn-Al2O3. In specific, at 260 °C, Pt/Al2O3 generated a rate constant of 0.00180 min−1, while Pt/Mg-Al2O3 and Pt/Zn-Al2O3 generated rate constants of 0.00140 and 0.00170 min−1, respectively. However, at temperatures > 280 °C, this changed drastically; the activity of the doped catalysts, particularly with Mg, increased from 0.0049 to 0.0137 min−1. The TOF followed a similar trend. A high TOF value indicates a high activity due to an increase in the number of available active sites for the reaction. A Mg-doped catalyst exhibited the highest activity, i.e., a TOF of 586 min−1, which is more than double compared to the Zn-doped catalyst (269 min−1) and almost three times than that obtained with an undoped catalyst (202 min−1) (calculated for dehydrogenation reactions at 300 °C).
The activation energy of the catalysts was calculated using the Arrhenius model, where the graph of ln k vs. 1/T is plotted (see Figure 11). The slope of the graph provided activation energies of 102, 130 and 151 kJ/mol for Pt/Al2O3, Pt/Zn-Al2O3 and Pt/Mg-Al2O3, respectively (see Table 4). A comparison indicates that our results are within the range reported in the literature. The regression was found to be >0.99 for all catalysts. Pt/Al2O3 had the lowest activation energy but low catalytic performance when compared to its counterparts (Pt/Zn-Al2O3 and Pt/Mg-Al2O3). This suggests that factors such as mass transfer limitations could have had an impact because a lower activation energy indicates that a reaction barrier is reduced and the reaction can proceed more rapidly. Interestingly, the pre-exponential factor, which indicates the frequency of the H18-DBT molecule encountered on the active sites, is directly proportional to the activation energy. The high pre-exponential factor associated with Pt/Mg-Al2O3 suggests an increase in the number of active sites. If the active sites are less effective (lower energy), then this decrease in energy of the active sites would lead to a growth in the activation energy [29]. The high activation energy observed for Pt/Mg-Al2O3 means more energy was used to overcome the reaction barrier when using this catalyst. Jorschick et al. argue that this high activation energy indicates that mass transfer effects have no dominant effect in the dehydrogenation reaction under the used conditions [30].
A similar trend was observed by Seidel [16]: the dehydrogenation of H18-DBT at 270–320 °C with 0.5 wt % Pt/Al2O3, 0.3 wt % Pt/Al2O3 and 0.3 wt % Pt/Al2O3 (0.14 wt % S) generated activation energies of 97, 131 and 143 kJ/mol, respectively. High activation energies (131 and 143 kJ/mol) for catalysts with large pore diameter (24.6 nm) were obtained (compared with 97 kJ/mol for a pore diameter 3.2 nm). Therefore, the addition of Mg and Zn dopants reduced the diffusion barrier, as evidenced by the low activation energy (102 kJ/mol) of the undoped catalyst. Mg also increases basic surface OH groups on catalysts, which enhances fast kinetics, as evidenced by the work of Kim et al. [31]. The Ru/MgO catalysts had better activity and kinetics compared with Ru/Al2O3 during the hydrogenation of benzyltoluene because of the low surface acidity of the MgO support. We are thus able to confirm the higher catalytic activity of the Mg-doped catalyst compared with the other catalysts considered here. Since the catalysts studied here had the same textural properties (see Table 2), the ability of the Mg catalyst to slightly suppress pore size reduction may have contributed to improved performance. Optimal Mg loading could further contribute to the retention of the pore size and subsequently result in better performance.
Table 4. Comparison of kinetic parameters of Pt/Al2O3 catalysts between this study and the literature.
Table 4. Comparison of kinetic parameters of Pt/Al2O3 catalysts between this study and the literature.
ReferenceReactor TypeCatalystConditionsKinetic Values
nk0Ea
This studyBatch0.5 wt % Pt/Al2O3260–290 °C; 0.2 mol %11.77 × 107102
0.5 wt % Pt/Mg-Al2O312.30 × 1011151
0.5 wt % Pt/Zn-Al2O311.15 × 1010131
[12]Batch1 wt % Pt/Al2O3260–290 °C; 0.4 mol %12.21 × 1016205
1 wt % Pd/Al2O312.26 × 10484
1 wt % Pt-Pd/Al2O312.35 × 10266
[18]Batch1 wt % Pt/Al2O3300–355 °C; 0.2 mol %11.76 × 1016201
[19]Fixed bed0.3 wt % Pt/Al2O3280–300 °C, 1 bar (2.5 bar)26.49 × 105 (2.66 × 108)117(149)
[15]Batch0.3 wt % Pt/TiO2270–310 °C; 0.1 mol %1.666.16 × 109145
0.3 wt % Pt (0.1 wt % S)/TiO2270–310 °C; 0.1 mol %1.513.23 × 1012174
[22]Fixed bedPt/Al2O3260–310 °C1125 s−1120
[20]Pd/Ag membranePt/Al2O3300–350 °C12.637 × 10−6 m3/kgcat.s156 ± 28.5
[32]Batch0.3 wt % Pt/Al2O3280–310 °C; 0.05 mol %; di 11 nm-1.60 × 1010151
0.3 wt % Pt/Al2O3280–310 °C; 0.05 mol %; di 14 nm-1.50 × 1011161
0.3 wt %Pt/Al2O3280–310 °C; 0.05 mol %; di 18 nm-1.30 × 1012169
[21]Batch0.5 wt % Pt/Al2O3270–320 °C; 0.05 mol %1.848.30 × 1097
0.3 wt % Pt/Al2O3270–320 °C; 0.05 mol %2.212.27 × 104131
0.3 wt % Pt (0.14 wt % S)/Al2O3270–320 °C; 0.05 mol %1.626.76 × 106143
[23]Fixed bed5 wt % Pt/Al2O3250–320 °C; WHSV > 2 h−12.35–2.445.40 × 1011171
[8]Microchannel reactor2 wt % Pt/Al2O3260–320 °C; 0.01 mL/min13.2725 s−113.79
n is the order of reaction, k0 is the frequency or pre-exponential factor in min−1 and Ea is the activation energy in kJ/mol.

3. Characterization

This study is a continuation of the previous work already published, where methods and detailed characterization of the catalysts used in this study can be found in the reported literature [13]. Characterization techniques, such as CO pulse chemisorption, Hydrogen temperature programmed reduction (H2-TPR), Ammonia temperature programmed desorption (NH3-TPD), transmission electron microscopy (TEM) and inductively coupled plasma-optical emission spectrometry (ICP-OES) are described in our previous study. The Pt dispersion percentage, a critical factor for catalytic efficiency, is highest in Pt/Al2O3 (38%), indicating a better spread of platinum particles across the support. However, the Mg-doped catalyst (34% dispersion) outperformed a highly dispersed Pt/Al2O3. Therefore, a 4% difference in dispersion has not led to a significant change in catalyst performance. Despite differences in dispersion, both Pt/Al2O3 and Pt/Mg-Al2O3 show identical metallic surface areas (0.42 m2/g), suggesting that the metallic surface area is not a determining factor for the improved performance shown by Pt/Mg-Al2O3. The hydrogen consumption values could also reflect the catalytic activity, which is notably higher for Pt/Mg-Al2O3 (64 µmol/g) and Pt/Zn-Al2O3 (55 µmol/g). High hydrogen consumption in TPR indicates the presence of a significant amount of reducible metal oxides, which in a metallic state are responsible for the catalytic activity. There is a slight variation in total acidity observed among the three catalysts, with marginally higher acidity in Pt/Zn-Al2O3. Particle size has an impact on the surface area and reaction kinetics, and Pt/Zn-Al2O3 exhibits slightly smaller particles with a 0.06–0.08 nm difference. The consistency in Pt loading across the catalysts is notable, but Mg and Zn dopants present in the latter two suggest a modification in electronic or structural properties, which can considerably affect catalytic performance already discussed. Thus, a comparison depicted in Table 5 emphasizes the importance of choosing a catalyst based on the specific requirements of the H18-DBT dehydrogenation process.

4. Experimental

The modified Pt-based catalysts (Pt/Al2O3, Pt/Mg-Al2O3, Pt/Zn-Al2O3) were prepared following the wet impregnation method as described, characterized and analyzed in our previous work [8]. A batch reactor setup, as shown in Figure 12, was used to study the reaction kinetics for the dehydrogenation of H18-DBT. A more detailed description of the experimental setup was reported earlier by our research team [16].
The degree of dehydrogenation (dod) obtained was calculated based on the total hydrogen produced vs. theoretical hydrogen stored in H18-DBT. This value was confirmed by determining the dod of the remaining reaction mixture using a calibrated refractometer. The dod, productivity (P) and concentration (C) were calculated using Equations (3), (4) and (5), respectively.
Dod   % = Volume   of   H 2   released Theoretical   H 2   volume × 100
P = m H 2 m Pt · t
C H 18 DBT = n H 18 DBT 1 Yield   ( % ) 100 V Total
where m, n and VTotal are the mass (g), number of moles and total volume (L) of the liquid, respectively.

5. Conclusions

The evaluation of the catalytic performance of Mg- and Zn-modified Pt/Al2O3 catalysts in a batch reactor indicated a strong dependence of activity and performance on temperature. The highest activities were obtained at 300 °C for all catalysts. The Pt/Mg-Al2O3 was the best-performing catalyst with dod, productivity, conversion and selectivity of 100%, 1.84 gH2/gPt/min, 99.9% and 87%, respectively. After run 4, the catalytic activity for all catalysts decreased due to catalyst deactivation. The extent of deactivation, expressed as a deactivation parameter, indicated that Mg reduced the deactivation of Pt/Al2O3 by a factor of three, which then led to improved stability and better performance. The presence of by-products was a function of catalyst activity. Mg-doped catalysts performed better because of their favourable acidity feature, which suppressed deactivation and facilitated easy diffusion of reactants and products (high activation term), thus resulting in better catalytic activity. The dehydrogenation activity of H18-DBT followed first-order reaction kinetics. The obtained activation energy of Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3 were 101, 151 and 131 kJ/mol, respectively. In this case, a high number of active sites corresponds to a high frequency factor value, which relates to high activation energy. Therefore, a lower activation energy for the dehydrogenation of H18-DBT will not always produce a high reaction rate. The high activation energy of the doped catalysts, particularly Pt/Mg-Al2O3, suggests that mass transfer effects have no dominant effect in the dehydrogenation reaction under the used conditions. The catalysts in stable form meet the set criteria in terms of H18-DBT conversion (>90%). However, there is a need for further improvement since the catalyst productivity and selectivity do not meet the set criteria provided in the introduction section.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14010032/s1, Figure S1: The effect of Mg and Zn dopants on the hydrogen flow and yield in a batch reactor. Reaction parameters as a function of dehydrogenation temperature in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt; nPt/nLOHC, 0.2 mol %; temperature, 300 °C; time, 90 min.

Author Contributions

Conceptualization, P.M. and D.B.; Methodology, R.G.; Formal analysis, R.G.; Data curation, R.G.; Writing—original draft, R.G. and P.M.; Supervision, P.M. and D.B.; Project administration, D.B.; Funding acquisition, D.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

Financial assistance from North West University the Department of Science and Innovation, South Africa, and Sasol South Africa is greatly appreciated.

Conflicts of Interest

The authors have no competing interests to declare that are relevant to the contents of this article.

References

  1. Preuster, P.; Fang, Q.; Peters, R.; Deja, R.; Nguyen, V.N.; Blum, L.; Stolten, D.; Wasserscheid, P. Solid Oxide Fuel Cell Operating on Liquid Organic Hydrogen Carrier-Based Hydrogen—Making Full Use of Heat Integration Potentials. Int. J. Hydrogen Energy 2018, 43, 1758–1768. [Google Scholar] [CrossRef]
  2. Fitchner, M.; Idrisova, F. Fundamental Properties of Hydrogen. In The Hydrogen Economy: Opportunities and Challenges; Cambridge University Press: New York, NY, USA, 2018; pp. 271–276. [Google Scholar]
  3. Aakko-Saksa, P.T.; Cook, C.; Kiviaho, J.; Repo, T. Liquid Organic Hydrogen Carriers for Transportation and Storing of Renewable Energy—Review and Discussion. J. Power Sources 2018, 396, 803–823. [Google Scholar] [CrossRef]
  4. Modisha, P.M.; Ouma, C.N.M.; Garidzirai, R.; Wasserscheid, P.; Bessarabov, D. The Prospect of Hydrogen Storage Using Liquid Organic Hydrogen Carriers. Energy Fuels 2019, 33, 2778–2796. [Google Scholar] [CrossRef]
  5. Brückner, N.; Obesser, K.; Bösmann, A.; Teichmann, D.; Arlt, W.; Dungs, J.; Wasserscheid, P. Evaluation of Industrially Applied Heat-Transfer Fluids as Liquid Organic Hydrogen Carrier Systems. ChemSusChem 2014, 7, 229–235. [Google Scholar] [CrossRef] [PubMed]
  6. Shi, L.; Zhou, Y.; Qi, S.; Smith, K.J.; Tan, X.; Yan, J.; Yi, C. Pt Catalysts Supported on H2 and O2 Plasma-Treated Al2O3 for Hydrogenation and Dehydrogenation of the Liquid Organic Hydrogen Carrier Pair Dibenzyltoluene and Perhydrodibenzyltoluene. ACS Catal. 2020, 10, 10661–10671. [Google Scholar] [CrossRef]
  7. Shi, L.; Qi, S.; Qu, J.; Che, T.; Yi, C.; Yang, B. Integration of Hydrogenation and Dehydrogenation Based on Dibenzyltoluene as Liquid Organic Hydrogen Energy Carrier. Int. J. Hydrogen Energy 2019, 44, 5345–5354. [Google Scholar] [CrossRef]
  8. Ali, A.; Rohini, A.K.; Lee, H.J. Dehydrogenation of Perhydro-Dibenzyltoluene for Hydrogen Production in a Microchannel Reactor. Int. J. Hydrogen Energy 2022, 47, 20905–20914. [Google Scholar] [CrossRef]
  9. Kwak, Y.; Kirk, J.; Moon, S.; Ohm, T.; Lee, Y.J.; Jang, M.; Park, L.H.; Il Ahn, C.; Jeong, H.; Sohn, H.; et al. Hydrogen Production from Homocyclic Liquid Organic Hydrogen Carriers (LOHCs): Benchmarking Studies and Energy-Economic Analyses. Energy Convers. Manag. 2021, 239, 114124. [Google Scholar] [CrossRef]
  10. Lee, S.; Han, G.; Kim, T.; Yoo, Y.S.; Jeon, S.Y.; Bae, J. Connected Evaluation of Polymer Electrolyte Membrane Fuel Cell with Dehydrogenation Reactor of Liquid Organic Hydrogen Carrier. Int. J. Hydrogen Energy 2020, 45, 13398–13405. [Google Scholar] [CrossRef]
  11. Lee, S.; Lee, J.; Kim, T.; Han, G.; Lee, J.; Lee, K.; Bae, J. Pt/CeO2 Catalyst Synthesized by Combustion Method for Dehydrogenation of Perhydro-Dibenzyltoluene as Liquid Organic Hydrogen Carrier: Effect of Pore Size and Metal Dispersion. Int. J. Hydrogen Energy 2021, 46, 5520–5529. [Google Scholar] [CrossRef]
  12. Modisha, P.; Gqogqa, P.; Garidzirai, R.; Ouma, C.N.M.; Bessarabov, D. Evaluation of Catalyst Activity for Release of Hydrogen from Liquid Organic Hydrogen Carriers. Int. J. Hydrogen Energy 2019, 44, 21926–21935. [Google Scholar] [CrossRef]
  13. Garidzirai, R.; Modisha, P.; Shuro, I.; Visagie, J.; van Helden, P.; Bessarabov, D. The Effect of Mg and Zn Dopants on Pt/Al2O3 for the Dehydrogenation of Perhydrodibenzyltoluene. Catalysts 2021, 11, 490. [Google Scholar] [CrossRef]
  14. Auer, F.; Blaumeiser, D.; Bauer, T.; Bösmann, A.; Szesni, N.; Libuda, J.; Wasserscheid, P. Boosting the Activity of Hydrogen Release from Liquid Organic Hydrogen Carrier Systems by Sulfur-Additives to Pt on Alumina Catalysts. Catal. Sci. Technol. 2019, 9, 3537–3547. [Google Scholar] [CrossRef]
  15. Chen, X.; Gierlich, C.H.; Schötz, S.; Blaumeiser, D.; Bauer, T.; Libuda, J.; Palkovits, R. Hydrogen Production Based on Liquid Organic Hydrogen Carriers through Sulfur Doped Platinum Catalysts Supported on TiO2. ACS Sustain. Chem. Eng. 2021, 9, 6561–6573. [Google Scholar] [CrossRef]
  16. Modisha, P.; Garidzirai, R.; Günes, H.; Bozbag, S.E.; Rommel, S.; Uzunlar, E.; Aindow, M.; Erkey, C.; Bessarabov, D. A Promising Catalyst for the Dehydrogenation of Perhydro-Dibenzyltoluene: Pt/Al2O3 Prepared by Supercritical CO2 Deposition. Catalysts 2022, 12, 489. [Google Scholar] [CrossRef]
  17. Catalyst Development for Improved Economic Viability of LOHC Technology. Available online: https://cordis.europa.eu/programme/id/H2020_FCH-02-1-2020 (accessed on 10 October 2023).
  18. Modisha, P.; Bessarabov, D. Stress Tolerance Assessment of Dibenzyltoluene-Based Liquid Organic Hydrogen Carriers. Sustain. Energy Fuels 2020, 4, 4662–4670. [Google Scholar] [CrossRef]
  19. Bulgarin, A.; Jorschick, H.; Preuster, P.; Bösmann, A.; Wasserscheid, P. Purity of Hydrogen Released from the Liquid Organic Hydrogen Carrier Compound Perhydro Dibenzyltoluene by Catalytic Dehydrogenation. Int. J. Hydrogen Energy 2020, 45, 712–720. [Google Scholar] [CrossRef]
  20. Wunsch, A.; Mohr, M.; Pfeifer, P. Intensified LOHC-Dehydrogenation Using Multi-Stage Microstructures and Pd-Based Membranes. Membranes 2018, 8, 112. [Google Scholar] [CrossRef]
  21. Seidel, A. Entwicklung Eines Technischen Platin-Trägerkatalysators zur Dehydrierung von Perhydro-Dibenzyltoluol. Ph.D. Thesis, Friedrich Alexander Universität Erlangen, Nürnberg, Germany, 2019; 177p. [Google Scholar]
  22. Peters, R.; Deja, R.; Fang, Q.; Nguyen, V.N.; Preuster, P.; Blum, L.; Wasserscheid, P.; Stolten, D. A Solid Oxide Fuel Cell Operating on Liquid Organic Hydrogen Carrier-Based Hydrogen—A Kinetic Model of the Hydrogen Release Unit and System Performance. Int. J. Hydrogen Energy 2019, 44, 13794–13806. [Google Scholar] [CrossRef]
  23. Park, S.; Naseem, M.; Lee, S. Experimental Assessment of Perhydro-Dibenzyltoluene Dehydrogenation Reaction Kinetics in a Continuous Flow System for Stable Hydrogen Supply. Materials 2021, 14, 7613. [Google Scholar] [CrossRef]
  24. Ballarini, A.; Basile, F.; Benito, P.; Bersani, I.; Fornasari, G.; De Miguel, S.; Maina, S.C.P.; Vilella, J.; Vaccari, A.; Scelza, O.A. Platinum Supported on Alkaline and Alkaline Earth Metal-Doped Alumina as Catalysts for Dry Reforming and Partial Oxidation of Methane. Appl. Catal. A Gen. 2012, 433–434, 1–11. [Google Scholar] [CrossRef]
  25. Belskaya, O.B.; Stepanova, L.N.; Gulyaeva, T.I.; Erenburg, S.B.; Trubina, S.V.; Kvashnina, K.; Nizovskii, A.I.; Kalinkin, A.V.; Zaikovskii, V.I.; Bukhtiyarov, V.I.; et al. Zinc Influence on the Formation and Properties of Pt/Mg(Zn)AlOx Catalysts Synthesized from Layered Hydroxides. J. Catal. 2016, 341, 13–23. [Google Scholar] [CrossRef]
  26. Jorschick, H.; Dürr, S.; Preuster, P.; Bösmann, A.; Wasserscheid, P. Operational Stability of a LOHC-Based Hot Pressure Swing Reactor for Hydrogen Storage. Energy Technol. 2019, 7, 146–152. [Google Scholar] [CrossRef]
  27. Modisha, P.M.; Jordaan, J.H.L.; Bösmann, A.; Wasserscheid, P.; Bessarabov, D. Analysis of Reaction Mixtures of Perhydro-Dibenzyltoluene Using Two-Dimensional Gas Chromatography and Single Quadrupole Gas Chromatography. Int. J. Hydrogen Energy 2018, 43, 5620–5636. [Google Scholar] [CrossRef]
  28. Sotoodeh, F. Hydrogenation and Dehydrogenation Kinetics and Catalysts for New Hydrogen Storage Liquids. Ph.D Thesis, University of British Columbia, Vancouver, BC, Canada, 2011; 223p. [Google Scholar]
  29. Triyono, T. Correlation between Preexponential Factor and Activation Energy of Isoamylalcohol Hydrogenolysis on Platinum Catalysts. Indones. J. Chem. 2010, 4, 1–5. [Google Scholar] [CrossRef]
  30. Jorschick, H.; Geißelbrecht, M.; Eßl, M.; Preuster, P.; Bösmann, A.; Wasserscheid, P. Benzyltoluene/Dibenzyltoluene-Based Mixtures as Suitable Liquid Organic Hydrogen Carrier Systems for Low Temperature Applications. Int. J. Hydrogen Energy 2020, 45, 14897–14906. [Google Scholar] [CrossRef]
  31. Kim, T.W.; Kim, M.; Kim, S.K.; Choi, Y.N.; Jung, M.; Oh, H.; Suh, Y.W. Remarkably Fast Low-Temperature Hydrogen Storage into Aromatic Benzyltoluenes over MgO-Supported Ru Nanoparticles with Homolytic and Heterolytic H2 Adsorption. Appl. Catal. B Environ. 2021, 286, 119889. [Google Scholar] [CrossRef]
  32. Auer, F. Katalysatorentwicklung für Die Dehydrierung von Perhydro-Dibenzyltoluol. Ph.D. Thesis, Friedrich Alexander Universität Erlangen, Nürnberg, Germany, 2020. [Google Scholar]
Figure 1. The effect of temperature on hydrogen flow, productivity and dod. Reaction parameters: H18-DBT degree of hydrogenation (doh), 99%; 0.5 wt % Pt/Al2O3; Pt to H18-DBT molar ratio (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; time, 90 min.
Figure 1. The effect of temperature on hydrogen flow, productivity and dod. Reaction parameters: H18-DBT degree of hydrogenation (doh), 99%; 0.5 wt % Pt/Al2O3; Pt to H18-DBT molar ratio (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; time, 90 min.
Catalysts 14 00032 g001
Figure 2. Comparison of productivity of doped and undoped Pt/Al2O3 catalysts used for H18-DBT dehydrogenation in a batch reactor: this study vs. literature [14,15,26,27].
Figure 2. Comparison of productivity of doped and undoped Pt/Al2O3 catalysts used for H18-DBT dehydrogenation in a batch reactor: this study vs. literature [14,15,26,27].
Catalysts 14 00032 g002
Figure 3. Total ion chromatogram of the LOHC reaction mixture obtained after 300 °C using Pt/Mg-Al2O3 catalyst. Insert: H18-DBT reaction pathway.
Figure 3. Total ion chromatogram of the LOHC reaction mixture obtained after 300 °C using Pt/Mg-Al2O3 catalyst. Insert: H18-DBT reaction pathway.
Catalysts 14 00032 g003
Figure 4. The H18-DBT conversion and H0-DBT selectivity vs. dehydrogenation temperature. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; reaction time, 90 min.
Figure 4. The H18-DBT conversion and H0-DBT selectivity vs. dehydrogenation temperature. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; reaction time, 90 min.
Catalysts 14 00032 g004
Figure 5. By-products formed during the dehydrogenation of H18-DBT in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; reaction time, 6 h.
Figure 5. By-products formed during the dehydrogenation of H18-DBT in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 260–300 °C; reaction time, 6 h.
Catalysts 14 00032 g005
Figure 6. The evolution of by-products and the degree of dehydrogenation as a function of time using Pt/Al2O3, Pt/Zn-Al2O3 and Pt/Mg-Al2O3. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt loading; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 300 °C.
Figure 6. The evolution of by-products and the degree of dehydrogenation as a function of time using Pt/Al2O3, Pt/Zn-Al2O3 and Pt/Mg-Al2O3. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt loading; molar ratio of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 300 °C.
Catalysts 14 00032 g006
Figure 7. Possible side reaction pathways for the formation of by-products.
Figure 7. Possible side reaction pathways for the formation of by-products.
Catalysts 14 00032 g007
Figure 8. Catalyst deactivation in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ration of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 300 °C; time, 6 h.
Figure 8. Catalyst deactivation in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt/Al2O3; molar ration of Pt to LOHC (nPt/nLOHC), 0.2 mol %; temperature, 300 °C; time, 6 h.
Catalysts 14 00032 g008
Figure 9. The change in concentration of H18-DBT versus time for Pt/Al2O3 catalysts in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt; nPt/nLOHC (molar ratio of Pt to H18-DBT), 0.2 mol %; temperature, 260–300 °C; reaction time, 90 min.
Figure 9. The change in concentration of H18-DBT versus time for Pt/Al2O3 catalysts in a batch reactor. Reaction parameters: H18-DBT, 99% doh; 0.5 wt % Pt; nPt/nLOHC (molar ratio of Pt to H18-DBT), 0.2 mol %; temperature, 260–300 °C; reaction time, 90 min.
Catalysts 14 00032 g009
Figure 10. First-order reaction kinetics of Pt/Al2O3 catalysts during H18-DBT dehydrogenation between 260 and 290 °C.
Figure 10. First-order reaction kinetics of Pt/Al2O3 catalysts during H18-DBT dehydrogenation between 260 and 290 °C.
Catalysts 14 00032 g010
Figure 11. Arrhenius plot for the dehydrogenation of H18-DBT between 260 and 290 °C.
Figure 11. Arrhenius plot for the dehydrogenation of H18-DBT between 260 and 290 °C.
Catalysts 14 00032 g011
Figure 12. Schematic representation of a batch reactor setup used for dehydrogenation experiments [6].
Figure 12. Schematic representation of a batch reactor setup used for dehydrogenation experiments [6].
Catalysts 14 00032 g012
Table 1. Quantities of by-products obtained for dehydrogenation of H18-DBT using different catalysts at 300 °C.
Table 1. Quantities of by-products obtained for dehydrogenation of H18-DBT using different catalysts at 300 °C.
By-Product Quantities in Mol. % per Catalyst Used
By-Product CompoundsCategoryPt/Al2O3Pt/Mg-Al2O3Pt/Zn-Al2O3
Benzyltoluene (m/z =182)Low boilers0.22.51.3
Toluene (m/z = 92)Low boilers0.010.30.2
H6-BT (m/z = 188)Low boilers0.21.20.6
Benzylmethylfluorene (m/z = 270)High boilers0.88.02.7
Total quantity (mol. %) 1.2125
Table 2. Summary of the effects of Mg and Zn dopants on Pt/Al2O3 for the H18-DBT dehydrogenation performance at 300 °C in a batch reactor.
Table 2. Summary of the effects of Mg and Zn dopants on Pt/Al2O3 for the H18-DBT dehydrogenation performance at 300 °C in a batch reactor.
CatalystDodiDodrun 4Pi
Prun 4
(gH2/gPt/min)
SiSrun 4XiXrun 4ΔX
Pt/Al2O382501.620.53793599.3927.4
Pt/Mg-Al2O3100731.840.58876299.9981.9
Pt/Zn-Al2O388601.680.55824999.9954.9
Xi and Xrun 4 are the initial and final H18-DBT conversions (%), Si and Srun 4 are the initial and final H0-DBT selectivities (%), Pi and Prun 4 are initial and final productivities, Yi and Y are the initial and final dod (%), X is the deactivation parameter (%), acidity is in mmol NH3/gcat.
Table 3. Rate constants, initial rates of reaction and TOF of H18-DBT dehydrogenation using Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3 catalysts between 260 and 300 °C.
Table 3. Rate constants, initial rates of reaction and TOF of H18-DBT dehydrogenation using Pt/Al2O3, Pt/Mg-Al2O3 and Pt/Zn-Al2O3 catalysts between 260 and 300 °C.
CatalystTemp, °Ck,
(min−1)
Initial Rate of Reaction
(M min−1 gcat−1)
TOF
(min−1)
Pt/Al2O32600.001800.0235247
2700.002700.0369670
2800.004200.05376109
2900.006000.08064155
3000.007800.09702202
Pt/Mg-Al2O32600.001400.0180661
2700.002400.03150103
2800.004900.06258209
2900.008300.09996355
3000.013700.17640586
Pt/Zn-Al2O32600.001700.0201649
2700.002800.0331881
2800.004900.05999142
2900.008000.09576231
3000.009300.10836269
Table 5. Catalyst properties determined by various characterization techniques.
Table 5. Catalyst properties determined by various characterization techniques.
Catalyst PropertiesPt/Al2O3Pt/Mg-Al2O3Pt/Zn-Al2O3
a Pt dispersion, %383423
a Metallic surface area, m2/gsample0.420.420.28
b H2 consumption, µmol/g346455
c Total Acidity, NH3/gcat0.480.410.49
d Particle size, nm1.721.701.54
e Pt loading, wt%0.480.490.48
e Dopant loading, wt % 3.893.44
a CO pulse chemisorption. b H2-TPR. c NH3-TPD. d TEM. e ICP-OES.
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

Garidzirai, R.; Modisha, P.; Bessarabov, D. Assessment of Reaction Kinetics for the Dehydrogenation of Perhydro-Dibenzyltoluene Using Mg- and Zn-Modified Pt/Al2O3 Catalysts. Catalysts 2024, 14, 32. https://doi.org/10.3390/catal14010032

AMA Style

Garidzirai R, Modisha P, Bessarabov D. Assessment of Reaction Kinetics for the Dehydrogenation of Perhydro-Dibenzyltoluene Using Mg- and Zn-Modified Pt/Al2O3 Catalysts. Catalysts. 2024; 14(1):32. https://doi.org/10.3390/catal14010032

Chicago/Turabian Style

Garidzirai, Rudaviro, Phillimon Modisha, and Dmitri Bessarabov. 2024. "Assessment of Reaction Kinetics for the Dehydrogenation of Perhydro-Dibenzyltoluene Using Mg- and Zn-Modified Pt/Al2O3 Catalysts" Catalysts 14, no. 1: 32. https://doi.org/10.3390/catal14010032

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