Next Article in Journal
The Effectiveness of Ni-Based Bimetallic Catalysts Supported by MgO-Modified Alumina in Dry Methane Reforming
Next Article in Special Issue
Construction of Spinel/Perovskite Heterojunction for Boosting Photocatalytic Performance for Polyacrylamide
Previous Article in Journal
Temperature-Dependent Hydrogenation, Hydrodeoxygenation, and Hydrogenolysis of Anisole on Nickel Catalysts
Previous Article in Special Issue
Cu-Doped SrTiO3 Nanostructured Catalysts for CO2 Conversion into Solar Fuels Using Localised Surface Plasmon Resonance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Propane Dehydrogenation over Cobalt Aluminates: Evaluation of Potential Catalytic Active Sites

by
Aleksey N. Chernov
,
Svetlana V. Cherepanova
,
Evgeny Yu. Gerasimov
,
Igor P. Prosvirin
,
Galina A. Zenkovets
,
Alexei A. Shutilov
,
Anna S. Gorbunova
,
Konstantin Yu. Koltunov
* and
Vladimir I. Sobolev
*
Boreskov Institute of Catalysis, Russian Academy of Sciences, Pr. Akademika Lavrentieva, 5, Novosibirsk 630090, Russia
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(11), 1419; https://doi.org/10.3390/catal13111419
Submission received: 12 October 2023 / Revised: 2 November 2023 / Accepted: 3 November 2023 / Published: 6 November 2023

Abstract

:
Non-oxidative propane dehydrogenation (PDH) is becoming an increasingly important approach to propylene production, while cobalt-containing catalysts have recently demonstrated great potential for use in this reaction, providing efficiencies comparable to those of industrially employed Pt- and Cr-based catalytic systems. It is therefore essential to clarify the nature of their active sites, especially since contradictory opinions on this issue are expressed in the literature. In this study, efforts were made to determine the state of Co in cobalt aluminates (CoAl2O4-Al2O3) responsible for PDH under typical operating conditions (600 °C, 1 atm). It is shown that the catalyst with a low cobalt content (Co/Al = 0.1) ensured the highest selectivity to propylene, ca. 95%, while maintaining significant propylene conversion. The structural motifs such as cobalt oxide and metallic cobalt nanoparticles, in addition to tetrahedral Co2+ species in the CoAl2O4 spinel system, were evaluated as potential active-site ensembles based on the obtained catalytic performance data in combination with the XRD, H2-TPR, TEM and XPS characteristics of as-synthesized, spent and spent–regenerated catalysts. It is revealed that the most likely catalytic sites linked to PDH are the Co-oxide forms tightly covering alumina or embedded in the spinel structure. However, additional in situ tuning is certainly needed, probably through the formation of surface oxygen vacancies rather than through a deeper reduction in Co0 as previously thought.

1. Introduction

Propylene, one of the basic organic chemicals, is largely produced by non-selective and energy-intensive naphtha steam cracking and fluid catalytic processes [1]. However, given the increased availability of propane as a feedstock, non-oxidative propane dehydrogenation (PDH) is now growing in importance as the most targeted and atom-efficient propylene production route [1,2,3,4,5,6,7,8,9,10]. Nonetheless, it is noteworthy that Pt- and CrOx-based catalysts commonly used in commercial PDH technologies have significant drawbacks, such as the high cost and difficult reactivation of the platinum catalyst and the negative environmental impact of chromium pollution [11].
Quite naturally, there is great interest in developing competitive PDH catalysts that do not contain precious metals and chromium. The most promising of them encompass a variety of transition metal oxides, namely CoOx, VOx, GaOx, MoOx, FeOx, WOx, InOx, ZnO, ZrO2 and TiO2, deposited on silica, alumina, zeolites and other oxide carriers [2,3,4,5,6,7,8,9,10,12,13]. The cobalt-containing catalysts should be especially highlighted because of their significant activity, relative affordability and moderate toxicity. In addition, carbon-based materials enriched with cobalt have also shown considerable potential in PDH [14,15,16,17,18,19,20]. However, remarkably, it is still not clear what structural motif in the cobalt catalysts promotes the PDH reaction. According to major conflicting opinions, Co+2 oxide forms, rather than metallic cobalt (Co0), and vice versa, may actually be the active sites [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43]. As a compromise, some authors attribute catalytic activity to both oxidation states, Co+2 and Co0 [33,35], or conclude that “the true active cobalt species for propane production are from tetrahedral Co+2 ions during the reaction” [21]. Additionally, apart from these possibilities, less attention has been drawn to the fact that low-valent first-row transition metal ions, including Co+1 in particular, can cause quite facile dehydrogenation of light alkanes [44,45,46,47]. Importantly, the oxidative addition/reductive elimination mechanism proposed for such reactions is independently well documented for a wide range of C–H bond functionalizations involving Co+1 organometallic compounds [48]. And lastly, atomically dispersed cobalt–nitrogen complexes deposited on carbon have also been suggested as possible catalytically active sites in the case of PDH over Co-N-C composites [14,15,16,17,18,19,20].
Given the strong interest in the activity of cobalt oxides in PDH, and cobalt–aluminum oxides in particular [21,22,23,26,28,29,33,34,41,43], we describe here a study on the behavior of CoAl2O4-Al2O3 catalysts in this reaction with the aim of elucidating the contributions of a variety of potential catalytic sites to catalytic performance.

2. Results

2.1. Catalyst Characterization

A number of textural characteristics of the CoAl2O4-Al2O3 catalysts, designated as CoAl-X (where X is the estimated molar ratio of Co/Al), are summarized in Table 1. The BET surface area was the highest for as-prepared CoAl-0.1 (291 m2 g−1), which is comparable to that of CoAl-0.05 (274 m2 g−1), despite the varying cobalt content, namely 6.7 and 4.3%, respectively. In addition, the total pore volume gradually increased as the cobalt content decreased. However, the BET surface areas and likewise the pore volumes of the spent–regenerated samples CoAl-0.25-R and CoAl-0.1-R were reduced markedly.
The XRD patterns of as-synthesized and spent–regenerated catalysts allowed us to estimate the relative contributions of cobalt oxide and aluminum oxide crystalline phases (Figure 1). Table S1 summarizes the results of refining the structure of these samples. As expected, the peaks of Co3O4 form for CoAl-0.5, CoAl-0.25 and spent–regenerated CoAl-0.25-R catalysts were much stronger than for other catalysts. It is noteworthy that the crystallinity level of Co3O4 in the CoAl-0.25-R sample increased, whereas in the case of the CoAl-0.1-R catalyst the opposite occurred (Table S1). In general, the XRD data show that only CoAl-0.05, CoAl-0.1 and CoAl-0.1-R samples demonstrated the clear domination of a mixed Co−Al oxide crystalline form, which can be related to a separate CoAl2O4 spinel structure.
Figure 2 shows the H2-TPR profiles of the catalysts, revealing that only CoAl-0.5, CoAl-0.25 and Co/Al2O3 exhibited a substantial H2 uptake. For these catalysts, the reduction temperature of the first peak attributed to the reduction of Co3O4 to CoO was about 500 °C, which is notably higher than that of bulk Co3O4 (200–400 °C [37,38]) due to the strong interaction of Co3O4 with Al2O3, which prevents the catalyst reduction. The second reduction peak is located between 550 and 750 °C. Overall, the two-step reduction process corresponds to sequential CoOx reduction into CoO and then into Co0 states of cobalt. The increased reduction temperature indicates that CoOx is tightly bound to the Al2O3 surface or forms a mixed oxide with it [37,38]. The weak hydrogen uptake for other catalysts (at least up to 600 °C and even higher) shows that their reduction is notably hindered, and the cobalt oxide phase remains largely resistant to reduction to a metal state.
According to TEM data (Figure 3), the CoAl-0.1 sample consists mainly of a separate phase of cobalt aluminate CoAl2O4 (Figure 3b). In addition, the catalyst contains Al2O3 crystals, which are uniformly covered by CoOx particles with a size of about 1–2 nm (Figure 3c). Notably, in Figure 3c the white particles are attributed to cobalt based on the relative mass (atomic number) of scattering Co and Al atoms, which affect the HAADF signal differently. In contrast, CoAl-0.25 contains, in addition to the above-mentioned species, well-distinguishable particles with an average grain size of 10–20 nm that can be assigned to both CoAl2O4 and Co3O4, taking into account their close lattice parameters (Figure 4).
To further characterize the nature of cobalt in the catalysts, the XPS spectra were obtained (Figure 5 and Table S2). In the high-resolution Co 2p spectra, the peaks centered at 781.3–781.8 (Co 2p3/2) and ~795 eV (Co 2p1/2) are related to Co2+ (Figure 5). An associated Co2+ satellite peak (shake-up peak) centered at ~787.0 eV is also present, confirming the predominant state of cobalt as tetrahedral Co2+ ions. As expected, the Al 2p and O 1s spectra of the catalysts are very similar, confirming the lack of differentiation in the electron state of aluminum and oxygen among the catalysts.

2.2. Gas-Phase Catalytic Reactions of Propane in a Flow Reactor

The activity of CoAl catalysts in PDH was estimated at 600 °C using a 10% propane/N2 mixture. Unlike other samples, CoAl-0.5 proved unsuitable to catalyze PDH, as it was prone to rapid carbonization, clogging the reactor. This applies to both the fresh and spent–regenerated catalysts, and can be explained by the presence of the relatively high content of an easily reduced massive CoOx form, which tends to induce cracking reactions [35,37,38,39,40]. In contrast, fresh CoAl-0.1 and CoAl-0.05 catalysts provided 47 and 36% of propane conversion after a 20 and 30 min induction period, respectively (Figure 6). Then, there was a gradual decrease in catalytic activity, while the selectivity to propylene remained at the initial level of ca. 95%. It is noteworthy that despite the two-and-a-half-times-higher Co content in the CoAl-0.25 catalyst (Table 1), its activity was nevertheless comparable to that of CoAl-0.1 (Figure 6). However, the induction period for CoAl-0.25 was twice as long, whereas selectivity to propylene was notably lower, reaching 90% only at 1.5 h. This result illustrates that the readily reducible extra-framework CoOx species still present in the fresh CoAl-0.25 catalyst impairs the target catalytic performance.
It should be noticed that under the specified reaction conditions, especially during the induction period, propane underwent concurrent cracking reactions yielding mainly methane, along with carbonaceous material (see below), which is typical for high-temperature PDH [2,3,4,5,6,7,8,9,10].
The pre-treatment of the fresh CoAl-0.1 catalyst with hydrogen (as described in Section 4.5) shortened the induction period and led to a slight increase in the maximal propane conversion (Figure 7). However, regeneration of the spent CoAl-0.1 catalyst by high-temperature treatment with oxygen (Section 4.5) reduced its activity by about 20% compared to the fresh catalyst (Figure 7a). On the other hand, the selectivity to propylene increased to 96% (Figure 7b). The same is true of the spent–regenerated CoAl-0.25-R catalyst, which provided notably higher selectivity to propylene, a shorter induction period, and well-reproducible catalytic activity tested in several runs (Figure 8).
Furthermore, the catalytic properties of the “reference” Co/Al2O3 sample turned out to be comparable with that of CoAl-0.05 (Figure S1). This is not surprising, taking into account the close content of cobalt (Table 1), while the other characteristics of these catalysts are somewhat different. The selectivity to propylene over Co/Al2O3 was even the highest among the tested catalysts, reaching 97% and 98% for fresh and spent–regenerated samples, respectively (Figure S2). However, the relative drop in maximal propane conversion (from 36 to 25%) after the catalyst regeneration was the most significant (Figure S2).

2.3. Structural Changes in Catalysts during PDH and Subsequent Regeneration

According to the literature, the deactivation of cobalt-containing catalysts during PDH is commonly due to the deposition of coke and the restructuring of cobalt species under high-temperature reducing conditions [19,20,28,29]. Indeed, the microscopic images of the spent CoAl-0.1 catalyst show considerable growth of carbon nanotubes (Figure 9a,b). However, encapsulation of Co-Al oxides in carbon did not occur (cf. [19,20]), which means that carbonization by itself is not a cause of deactivation. More important is the fact of partial reduction of cobalt to the metallic state with the formation of Co nanoparticles of relatively large sizes about 10 nm (Figure 9b,d,e), although small particles of Co on alumina did not undergo substantial aggregation (Figure 9c).
Regeneration of the CoAl-0.1 spent catalyst by high-temperature treatment with molecular oxygen results in the redispersion of aggregated cobalt nanoparticles, providing a catalyst quite similar in appearance to the original (Figure 10). The CoAl-0.25 catalyst also restores its structure after the use in PDH and subsequent heat treatment in oxygen, although more significant differentiation in CoAl-0.25-R morphology has occurred. Thus, separate phases of alumina (weakly enriched with cobalt), cobalt aluminate and rather large particles of Co3O4 can be observed via TEM (Figures S3 and S4). Remarkably, according to H2-TPR data, the CoOx species in CoAl-0.25-R are not so readily reducible with hydrogen, as was found for CoOx in the fresh CoAl-0.25 catalyst (Figure 2).

3. Discussion

The CoAl-0.1 sample, which demonstrated the best catalytic performance in PDH, comprises three structural elements: Al2O3, small CoOx particles (in the size range of 1–2 nm) on its surface and adjacent areas of crystalline cobalt aluminate (CoAl2O4). The CoAl-0.25 catalyst, exhibiting comparable catalytic activity, contains, in addition, Co3O4 particles of about 10 nm size and the Co3O4 phase of a comparatively low crystallinity (XRD data, Figure 1). The latter phase was easily reduced and could be responsible for the lower selectivity to propylene and longer induction period compared to the CoAl-0.1 catalyst. Indeed, CoAl-0.25-R, showing better selectivity to propylene and a shortened induction period (Figure 8), differed from CoAl-0.25 through its increased Co3O4 particle size (up to 100 nm, Figure S4) and their increased crystallinity (Figure 1, Table S1). Additionally, CoAl-0.25-R became practically non-reducible with hydrogen until reaching 700 °C (Figure 2). Furthermore, the high-temperature treatment with oxygen of both of the spent CoAl-0.1 and CoAl-0.25 catalysts resulted in a marked decrease in specific surface area (Table 1), which may be one of the reasons for the partial loss of their activity.
It is obvious that the target activity of CoAl catalysts in PDH may be associated with the small CoOx particles on alumina, as well as with comparatively large and slowly reducible CoAl2O4 and Co3O4 crystalline phases, which can undergo a common evolutionary chain during PDH or pre-reduction with hydrogen: Co+2/+3 → framework Co+2 → isolated Co0 → small-size (Co0-Co0)n cluster → large cobalt particle (cf. [35,36,37,38,39]). Among these forms, as mentioned in Section 1, Co+2 oxides and/or small Co0 species are widely suggested to be the active catalytic sites [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43].
It is noteworthy that the basic rationale for the key contribution of tetrahedral Co+2 ions is their exclusive observation by physical–chemical techniques with no detection of Co0 under PDH conditions [21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36]. The proposed reaction mechanism suggests a non-redox pathway involving the heterolytic C–H bond cleavage over four-coordinated Co+2 incorporated in the framework of oxide support [25,36]. In contrast, extra-framework Co3O4 weakly bonded to the support (if present) can easily be reduced to comparatively large metallic cobalt particles [35,37,38,39,40], leading to propane cracking until these zones are strongly carbonized [19,20,21,22,33].
However, equally convincing is the evidence that well-dispersed Co0 species generated in situ by partial reduction of Co+2 ions are the alternative active catalytic sites [33,35,40,41,42]. The main argument here is based on the apparent relationship between the appearance of cobalt nanoparticles and the increase in catalytic activity. Furthermore, Kondratenko et al. have shown recently that propylene formation depends linearly on Co0 accumulation in Co/silicalite-1 catalysts [43]. Though perhaps, the catalytic activity could simply correlate with the steady reduction of Co+2 to Co0 due to a related but different process.
With regard to the CoAl catalytic system, it should be noted that the most effective PDH catalysts, such as CoAl-0.05, CoAl-0.1, CoAl-0.1-R and CoAl-0.25-R, are actually most resistant to reduction (metallization) by hydrogen (Figure 2). Moreover, a slow reduction of CoOx to metallic cobalt followed by its aggregation to cobalt nanoparticles (Figure 9b) is accompanied by the catalyst’s progressive deactivation, which may be related. On the other hand, the key catalytic activity of four-coordinated Co+2 ions only (whether they are part of CoOx particles on aluminum or are included in CoAl2O4 and Co3O4 crystals) cannot explain the results of numerous case studies where a definite relationship between of the catalytic activity and Co0 accumulation was found. Nevertheless, this apparent contradiction can be resolved by examining the missing link between the cobalt states Co+2 and Co0 under the PDH reaction.
Indeed, a close literary analogy of the process of nickel(II) oxide reduction by hydrogen reveals that NiO(100) crystals exhibiting negligible reactivity to H2 become active in splitting H–H bonds when surface defects, namely O vacancies (OVs), are created [49,50,51]. Remarkably, a correlation was found between the concentration of OVs and the rate of NiO reduction to Ni0 [49]. It seems reasonable that a similar process of OV generation occurs on the surface of tetrahedral Co+2 oxide forms in the course of PDH. The lattice oxygen consumption leading to the formation of OVs under conditions of thermal gaseous reduction is quite typical for transition metal oxides, including CoOx [52,53,54,55]. Moreover, the importance of OVs for efficient C−H bond activation in PDH has already been proven for a number of metal oxides, such as CeO2 [56] and TiO2 [57,58], and even for “non-reducible” oxides, such as ZrO2 [59] and Al2O3 [60]. Likewise, the catalytic activity of WO3 in PDH, provided by pre-treatment with hydrogen, may have an analogous origin [61].
The involvement of OVs could explain the induction period normally required for PDH catalyzed by cobalt oxides, the positive effect of pre-reduction with hydrogen (see, for example, Figure 7) and the observed correlation of catalytic activity with Co0 accumulation, while an overly extensive reduction of Co+2 to Co0 causes catalyst deactivation. Also notably, Co atoms surrounding OVs and carrying uncompensated positive charge can actually simulate Co+1 species able to catalyze PDH through the oxidative addition/reductive elimination pathway (Scheme 1), thus linking together the Co+2, Co+1 and Co0 forms, considered earlier as alternatives in the literature.
Summarizing the above considerations, nanocrystalline CoAl2O4 and Co3O4, as well as fine CoOx particles on alumina (all present in CoAl active catalysts) may exhibit the same catalytic properties due to similar evolution under PDH conditions, namely partial reduction to metallic cobalt (ultimately) through the intermediacy of surface OVs. The small content of cobalt in CoAl catalysts ensures their strong interaction with alumina and embedding in the spinel structure, which generally prevents deep reduction of CoOx to Co0 under PDH conditions and provides the catalyst’s relative stability. Based on these assumptions, progress in the design of cobalt-containing catalysts and similar transition metal oxide catalytic systems will require their optimal ability to generate surface Ovs, along with their resistance to subsequent reduction and metallization to be taken into account.

4. Materials and Methods

4.1. Materials

Cobalt(II) nitrate hexahydrate (Co(NO3)2·6H2O, ≥98%, Fisher Scientific UK, Loughborough, UK) and aluminum nitrate nonahydrate (Al(NO3)3·9H2O, ≥98%, Sigma-Aldrich, Merck Life Science LLC, Moscow, Russia) were used as received from the chemical suppliers.

4.2. Catalyst Preparation

The cobalt aluminate spinels (CoAl2O4-Al2O3) were obtained by mixing 2 M Co(NO3)2 and 2 M Al(NO3)3 aqueous solutions followed by co-precipitating with a 12.5% NH4OH at constant pH 7.5 and temperature of 70 °C. The resulting precipitates were filtered and washed with distilled water until the ammonia was removed. The obtained materials were dried to an air-dry state at room temperature, then at 110 °C for 12 h, and calcined in air at 650 °C for 4 h in a muffle furnace. Accordingly, based on the molar ratio of cobalt and aluminum salts taken, the catalyst samples, denoted as CoAl-0.5, CoAl-0.25, CoAl-0.1 and CoAl-0.05, were prepared in the amount of 50 g each (Table 1).
The Co/Al2O3 catalyst was prepared via the incipient wetness impregnation of γ-Al2O3 with a cobalt(II) nitrate solution followed by drying at 110 °C for 12 h and thermal treatment in air at 650 °C for 4 h.

4.3. Catalyst Characterization Techniques

The X-ray diffraction (XRD) data were obtained on the diffractometer ARL X’tra (Thermo Fisher Scientific, Waltham, MA, USA) using monochromatic CuKα radiation (λ = 1.5418 Å) in a range of 2θ = 15−80° with a step of 0.02° (1°/min) using linear detector Mythen2R 1D (Dectris, Baden-Daettwil, Switzerland) with an angle of 7.0°. Refinement of the lattice parameters and determination of average crystallite sizes and weight ratios of Co3O4/Al2O3 were carried out by the Rietveld method applied for a range of 2θ = 63−70° using the TOPAS 4.2 program and structural data from the ICSD database (Table S1).
Temperature-programmed reduction (H2-TPR) was performed in a flow system with a thermal conductivity detector. The samples (0.05 g, 0.25–0.5 mm fraction) were heated at rates of 10 °C min−1 to 900 °C; the flow rate of the reducing mixture (10% H2 in argon) was 40 cm3 min−1.
The porous structure parameters were determined from N2 adsorption isotherms at 77 K on an ASAP-2400 (Micromeritics, Norcross, GA, USA) analyzer after degassing the samples at 150 °C to a residual pressure of 30 mTorr (4 Pa).
The content of Co element in the samples was measured with an Optima 4300 DV inductively coupled plasma optical emission spectrometer (ICP-OES) (PerkinElmer, Wellesley, MA, USA).
High-resolution transmission electron microscopy (HRTEM) and scanning transmission electron microscopy (STEM) were conducted with a Themis Z (Thermo Fisher Scientific, Waltham, MA, USA) instrument with an accelerating voltage of 200 kV and lattice resolution of 0.07 nm. The TEM images were recorded with a Ceta 16 (Thermo Fisher Scientific, USA) CCD matrix. Elemental analysis was performed with a Super-X EDS detector (Thermo Fisher Scientific, USA). High-angle annular dark-field (HAADF) STEM images were recorded using a standard ThemisZ detector.
X-ray photoelectron spectra (XPS) were recorded on a SPECS (Berlin, Germany) photoelectron spectrometer using a hemispherical PHOIBOS-150-MCD-9 analyzer (Al Kα radiation, hν = 1486.6 eV, 150 W). The binding energy (BE) scale was pre-calibrated using the positions of the peaks of Au 4f7/2 (BE = 84.0 eV) and Cu 2p3/2 (BE = 932.67 eV) core levels. The BE values were corrected by reference to the C 1s internal standard (284.5 eV).

4.4. Catalytic Performance Tests

The PDH reactions were carried out at atmospheric pressure in a 7.5 mm quartz tube reactor loaded with a catalyst (0.2 g, 0.25–0.5 mm fraction), mixed with 1 mL quartz particles (d = 0.45 to 1 mm). The gas mixture (10 vol% propane balanced with nitrogen) was fed to the reactor at a gas hourly space velocity (GHSV) of 7500 mL h−1 gcat−1 (propane weight hourly space velocity of 1.4 h−1). The reaction temperature, 600 °C, was controlled inside the reactor with an accuracy of ±0.5 °C. The reaction products were analyzed by an online gas chromatograph (Chromos GC-1000, Chromos Engineering, Dzerzhinsk, Russia) with a flame ionization detector, using a quartz capillary column of an inner diameter of 0.32 mm and a length of 7 m. The column was prepared by the static high-pressure method, yielding a monolithic layer of porous SiO2 formed in the process of sol-gel synthesis directly inside the capillary [62]. Prior to the measurements, the catalysts were heated at 600 °C in a flow of N2 for 1 h. The blank experiment (the Al2O3/ quartz particles only) showed propane conversion of ~1.2% and selectivity to propylene of about 70%.
The propane conversion (X) and propylene selectivity (S) were calculated according to the carbon atom balance method:
X C 3 H 8 = C C 3 H 8 0 C C 3 H 8 C C 3 H 8 0 × 100 % ;
S C 3 H 6 = C C 3 H 6 C C 3 H 8 0 C C 3 H 8 × 100 % ;
where C C 3 H 8 and C C 3 H 6 are the mole flow rates of propane and propylene, respectively. In view of the reaction-induced changes in the gas flow due to the conversion of one molecule of propane into two molecules (propylene and hydrogen) during the reaction, the values of C C 3 H 8 and C C 3 H 6 were normalized by carbon balance accordingly.

4.5. Catalyst Pre-Reduction with H2 and Regeneration Procedures

The treatment of CoAl catalysts with hydrogen was carried out in the reactor at 600 °C in the H2 flow (15 mL min−1) for 1.5 h and then in the N2 flow (30 mL min−1) for 1 h.
After catalytic testing in PDH, the spent CoAl catalysts were treated in the reactor at 600 °C in the O2 flow (15 mL min−1) for 1.5 h and then in the N2 flow (30 mL min−1) for 1 h.

5. Conclusions

It was shown that cobalt aluminate containing 4–7 wt.% Co, and as a consequence, consisting mainly of alumina (with inclusions of small CoOx particles on the surface) and CoAl2O4 spinel crystallites, can be an effective catalyst for PDH reactions under practical operating conditions (600 °C, 1 atm). Further increasing the Co content to 17 wt.% resulted in the formation of an additional Co3O4 phase without increasing the catalyst efficiency towards PDH.
The catalysts’ loss of activity with time on stream is related to cobalt aggregation in the form of metallic Co nanoparticles; however, coke deposition, largely in the form of carbon nanotubes, does not appear to be the critical reason for the catalysts’ deactivation. The high-temperature treatment of spent catalysts with molecular oxygen leads to their regeneration with a partial decrease in activity, but at the same time is accompanied by the stabilization of the catalysts’ operation.
The structural motifs, such as tetrahedral Co2+ species located in the CoAl2O4 and Co3O4 spinel forms, as well as in the small CoOx particles tightly covering alumina, can be regarded as the feasible primary active-site ensembles. However, their subsequent evolution under PDH reaction conditions, consisting of a partial reduction to metallic cobalt through the intermediacy of surface OVs, may provide further rationale for the origin of the catalytic activity of cobalt aluminates in PDH.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13111419/s1, Table S1: Grid parameters (a), average crystallite sizes and weight ratios; Figure S1: Propane conversion and selectivity to propylene as functions of time on stream over CoAl and Co/Al2O3 catalysts; Figure S2: Propane conversion and selectivity to propylene as functions of time on stream over Co/Al2O3 and Co/Al2O3-R catalysts; Figure S3: Low- and higher-magnification TEM images of the spent–regenerated CoAl-0.25-R catalyst; Figure S4: HAADF-STEM and related EDX-STEM Co and Al elemental mapping images of different areas of CoAl-0.25-R; Table S2: High-resolution XPS spectra of Al 2p and O 1s of the samples; Table S3: The N2 adsorption/desorption isotherms and the pore size distribution of the samples.

Author Contributions

Conceptualization, A.N.C.; investigation, A.N.C., S.V.C., E.Y.G., I.P.P., G.A.Z., A.S.G. and A.A.S.; data curation, K.Y.K. and V.I.S.; writing—original draft preparation, K.Y.K.; writing—review and editing, K.Y.K. and V.I.S.; supervision, V.I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education of the Russian Federation within the governmental order for the Boreskov Institute of Catalysis (project AAAA-A21-121011390008-4).

Data Availability Statement

Not applicable.

Acknowledgments

The studies were carried out using the facilities of the shared research center “National center of investigation of catalysts” at Boreskov Institute of Catalysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zimmermann, H. Propene. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co., KGaA: Weinheim, Germany, 2013. [Google Scholar] [CrossRef]
  2. Sattler, J.J.H.B.; Ruiz-Martinez, J.; Santillan-Jimenez, E.; Weckhuysen, B.M. Catalytic dehydrogenation of light alkanes on metals and metal oxides. Chem. Rev. 2014, 114, 10613–10653. [Google Scholar] [CrossRef] [PubMed]
  3. Hu, Z.-P.; Yang, D.; Wang, Z.; Yuan, Z.-Y. State-of-the-art catalysts for direct dehydrogenation of propane to propylene. Chin. J. Catal. 2019, 40, 1233–1254. [Google Scholar] [CrossRef]
  4. Martino, M.; Meloni, E.; Festa, G.; Palma, V. Propylene Synthesis: Recent Advances in the Use of Pt-Based Catalysts for Propane Dehydrogenation Reaction. Catalysts 2021, 11, 1070. [Google Scholar] [CrossRef]
  5. Dong, S.; Altvater, N.R.; Mark, L.O.; Hermans, I. Assessment and comparison of ordered & non-ordered supported metal oxide catalysts for upgrading propane to propylene. Appl. Catal. A-Gen. 2021, 617, 118121. [Google Scholar] [CrossRef]
  6. Chen, S.; Chang, X.; Sun, G.; Zhang, T.; Xu, Y.; Wang, Y.; Pei, C.; Gong, J. Propane dehydrogenation: Catalyst development, new chemistry, and emerging technologies. Chem. Soc. Rev. 2021, 50, 3315–3354. [Google Scholar] [CrossRef] [PubMed]
  7. Otroshchenko, T.; Jiang, G.; Kondratenko, V.A.; Rodemerck, U.; Kondratenko, E.V. Current status and perspectives in oxidative, non-oxidative and CO2-mediated dehydrogenation of propane and isobutane over metal oxide catalysts. Chem. Soc. Rev. 2021, 50, 473–527. [Google Scholar] [CrossRef]
  8. Dai, Y.; Gao, X.; Wang, Q.; Wan, X.; Zhou, C.; Yang, Y. Recent progress in heterogeneous metal and metal oxide catalysts for direct dehydrogenation of ethane and propane. Chem. Soc. Rev. 2021, 50, 5590–5630. [Google Scholar] [CrossRef]
  9. Wang, Y.; Hu, P.; Yang, J.; Zhu, Y.-A.; Chen, D. C–H bond activation in light alkanes: A theoretical perspective. Chem. Soc. Rev. 2021, 50, 4299–4358. [Google Scholar] [CrossRef]
  10. Liu, S.; Zhang, B.; Liu, G. Metal-based catalysts for the non-oxidative dehydrogenation of light alkanes to light olefins. React. Chem. Eng. 2021, 6, 9–26. [Google Scholar] [CrossRef]
  11. Phung, T.K.; Pham, T.L.M.; Vu, K.B.; Busca, G. (Bio)Propylene production processes: A critical review. J. Environ. Chem. Eng. 2021, 9, 105673. [Google Scholar] [CrossRef]
  12. Yuan, Y.; Lee, J.S.; Lobo, R.F. Ga+-Chabazite Zeolite: A Highly Selective Catalyst for Nonoxidative Propane Dehydrogenation. J. Am. Chem. Soc. 2022, 144, 15079–15092. [Google Scholar] [CrossRef]
  13. Yuan, Y.; Lobo, R.F. Zinc Speciation and Propane Dehydrogenation in Zn/H-ZSM-5 Catalysts. ACS Catal. 2023, 13, 4971–4984. [Google Scholar] [CrossRef]
  14. Xie, J.; Kammert, J.D.; Kaylor, N.; Zheng, J.W.; Choi, E.; Pham, H.N.; Sang, X.; Stavitski, E.; Attenkofer, K.; Unocic, R.R.; et al. Atomically Dispersed Co and Cu on N-Doped Carbon for Reactions Involving C–H Activation. ACS Catal. 2018, 8, 3875–3884. [Google Scholar] [CrossRef]
  15. Cao, T.; Dai, X.; Li, F.; Liu, W.; Bai, Y.; Fu, Y.; Qi, W. Efficient Non-Precious Metal Catalyst for Propane Dehydrogenation: Atomically Dispersed Cobalt-nitrogen Compounds on Carbon Nanotubes. ChemCatChem 2021, 13, 3067–3073. [Google Scholar] [CrossRef]
  16. Wang, Y.; Suo, Y.; Ren, J.-T.; Wang, Z.; Yuan, Z.-Y. Spatially isolated cobalt oxide sites derived from MOFs for direct propane dehydrogenation. J. Colloid Interface Sci. 2021, 594, 113–121. [Google Scholar] [CrossRef]
  17. Wang, Q.; Xu, W.; Ma, Z.; Yu, F.; Chen, Y.; Liao, H.; Wang, X.; Zhou, J. Highly Effective Direct Dehydrogenation of Propane to Propylene by Microwave Catalysis at Low Temperature over Co−Sn/NC Microwave Catalyst. ChemCatChem 2021, 13, 1009–1022. [Google Scholar] [CrossRef]
  18. Li, Y.-M.; Liu, Z.-Y.; Zhang, Q.-Y.; Wang, Y.-J.; Cui, G.-Q.; Zhao, Z.; Xu, C.-M.; Jiang, G.-Y. Influence of carbonization temperature on cobalt-based nitrogen-doped carbon nanopolyhedra derived from ZIF-67 for nonoxidative propane dehydrogenation. Pet. Sci. 2022, 20, 559–568. [Google Scholar] [CrossRef]
  19. Chernov, A.N.; Sobolev, V.I.; Koltunov, K.Y. Propane dehydrogenation to propylene over Co@N-doped carbon: Structure-activity-selectivity relationships. Catal. Commun. 2022, 170, 106495. [Google Scholar] [CrossRef]
  20. Chernov, A.N.; Sobolev, V.I.; Gerasimov, E.Y.; Koltunov, K.Y. Propane Dehydrogenation on Co-NC/SiO2 Catalyst: The Role of Single-Atom Active Sites. Catalysts 2022, 12, 1262. [Google Scholar] [CrossRef]
  21. Sun, Y.; Wu, Y.; Shan, H.; Li, C. Studies on the Nature of Active Cobalt Species for the Production of Methane and Propylene in Catalytic Dehydrogenation of Propane. Catal. Lett. 2015, 145, 1413–1419. [Google Scholar] [CrossRef]
  22. Sun, Y.N.; Gao, Y.N.; Wu, Y.M.; Shan, H.H.; Wang, G.W.; Li, C.Y. Effect of sulfate addition on the performance of Co/Al2O3 catalysts in catalytic dehydrogenation of propane. Catal. Commun. 2015, 60, 42–45. [Google Scholar] [CrossRef]
  23. Li, W.; Yu, S.Y.; Meitzner, G.D.; Iglesia, E. Structure and Properties of Cobalt-Exchanged H-ZSM5 Catalysts for Dehydrogenation and Dehydrocyclization of Alkanes. J. Phys. Chem. B 2001, 105, 1176–1184. [Google Scholar] [CrossRef]
  24. Hu, B.; Getsoian, A.; Schweitzer, N.M.; Das, U.; Kim, H.; Niklas, J.; Poluektov, O.; Curtiss, L.A.; Stair, P.C.; Miller, J.T.; et al. Selective propane dehydrogenation with single-site Co-II on SiO2 by a non-redox mechanism. J. Catal. 2015, 322, 24–37. [Google Scholar] [CrossRef]
  25. Estes, D.P.; Siddiqi, G.; Allouche, F.; Kovtunov, K.V.; Safonova, O.V.; Trigub, A.L.; Koptyug, I.V.; Coperet, C. C−H Activation on Co,O Sites: Isolated Surface Sites versus Molecular Analogs. J. Am. Chem. Soc. 2016, 138, 14987–14997. [Google Scholar] [CrossRef] [PubMed]
  26. Hu, B.; Kim, W.-G.; Sulmonetti, T.P.; Sarazen, M.L.; Tan, S.; So, J.; Liu, Y.; Dixit, R.S.; Nair, S.; Jones, C.W. A Mesoporous Cobalt Aluminate Spinel Catalyst for Nonoxidative Propane Dehydrogenation. ChemCatChem 2017, 9, 3330–3337. [Google Scholar] [CrossRef]
  27. Song, S.J.; Li, J.; Wu, Z.J.; Zhang, P.; Sun, Y.Q.; Song, W.Y.; Li, Z.X.; Liu, J. In situ encapsulated subnanometric CoO clusters within silicalite-1 zeolite for efficient propane dehydrogenation. AIChE J. 2022, 68, e17451. [Google Scholar] [CrossRef]
  28. Dewangan, N.; Ashok, J.; Sethia, M.; Das, S.; Pati, S.; Kus, H.; Kawi, S. Cobalt-Based Catalyst Supported on Different Morphologies of Alumina for Non-oxidative Propane Dehydrogenation: Effect of Metal Support Interaction and Lewis Acidic Sites. ChemCatChem 2019, 11, 4923–4934. [Google Scholar] [CrossRef]
  29. Dai, Y.; Gu, J.; Tian, S.; Wu, Y.; Chen, J.; Li, F.; Du, Y.; Peng, L.; Ding, W.; Yang, Y. γ-Al2O3 sheet-stabilized isolate Co2+ for catalytic propane dehydrogenation. J. Catal. 2020, 381, 482–492. [Google Scholar] [CrossRef]
  30. Bian, Z.F.; Dewangan, N.; Wang, Z.G.; Pati, S.; Xi, S.B.; Borgna, A.; Kus, H.; Kawi, S. Mesoporous-Silica-Stabilized Cobalt(II) Oxide Nanoclusters for Propane Dehydrogenation. ACS Appl. Nano Mater. 2021, 4, 1112–1125. [Google Scholar] [CrossRef]
  31. Wu, L.Z.; Ren, Z.Z.; He, Y.S.; Yang, M.; Yu, Y.K.; Liu, Y.M.; Tan, L.; Tang, Y. Atomically Dispersed Co2+ Sites Incorporated into a Silicalite-1 Zeolite Framework as a High Performance and Coking-Resistant Catalyst for Propane Nonoxidative Dehydrogenation to Propylene. ACS Appl. Mater. Interfaces 2021, 13, 48934–48948. [Google Scholar] [CrossRef]
  32. Ren, Z.; He, Y.; Yang, M.; Deng, H.; Zhang, Y.; Yang, H.; Tang, Z.; Tan, L.; Tang, Y.; Wu, L. The investigation into the different Co species over Silicalite-1 via modulating heat-treatment atmosphere for propane dehydrogenation. Mol. Catal. 2022, 530, 112580. [Google Scholar] [CrossRef]
  33. Jeon, N.; Oh, J.; Tayal, A.; Jeong, B.; Seo, O.; Kim, S.; Chung, I.; Yun, Y.J. Effects of heat-treatment atmosphere and temperature on cobalt species in Co/Al2O3 catalyst for propane dehydrogenation. J. Catal. 2021, 404, 1007–1016. [Google Scholar] [CrossRef]
  34. Ge, M.; Chen, X.; Li, Y.; Wang, J.; Xu, Y.; Zhang, L. Perovskite-derived cobalt-based catalyst for catalytic propane dehydrogenation. React. Kinet. Mech. Catal. 2020, 130, 241–256. [Google Scholar] [CrossRef]
  35. Gao, Y.; Peng, L.; Long, J.; Wu, Y.; Dai, Y.; Yang, Y. Hydrogen pre-reduction determined Co–silica interaction and performance of cobalt catalysts for propane dehydrogenation. Micropor. Mesopor. Mater. 2021, 323, 111187. [Google Scholar] [CrossRef]
  36. Hu, Z.-P.; Qin, G.; Han, J.; Zhang, W.; Wang, N.; Zheng, Y.; Jiang, Q.; Ji, T.; Yuan, Z.-Y.; Xiao, J.; et al. Atomic Insight into the Local Structure and Microenvironment of Isolated Co-Motifs in MFI Zeolite Frameworks for Propane Dehydrogenation. J. Am. Chem. Soc. 2022, 144, 12127–12137. [Google Scholar] [CrossRef]
  37. Bulavchenko, O.A.; Cherepanova, S.V.; Malakhov, V.V.; Dovlitova, L.S.; Ishchenko, A.V.; Tsybulya, S.V. In situ XRD study of nanocrystalline cobalt oxide reduction. Kinet. Catal. 2009, 50, 192–198. [Google Scholar] [CrossRef]
  38. Ji, Y.; Zhao, Z.; Duan, A.; Jiang, G.; Liu, J. Comparative Study on the Formation and Reduction of Bulk and Al2O3-Supported Cobalt Oxides by H2-TPR Technique. J. Phys. Chem. C 2009, 113, 7186–7199. [Google Scholar] [CrossRef]
  39. Tomic-Tucakovic, B.; Majstorovic, D.; Jelic, D.; Mentus, S. Thermogravimetric study of the kinetics of Co3O4 reduction by hydrogen. Thermochim. Acta 2012, 541, 15–24. [Google Scholar] [CrossRef]
  40. Chen, C.; Zhang, S.M.; Wang, Z.; Yuan, Z.Y. Ultrasmall Co confined in the silanols of dealuminated beta zeolite: A highly active and selective catalyst for direct dehydrogenation of propane to propylene. J. Catal. 2020, 383, 77–87. [Google Scholar] [CrossRef]
  41. Li, X.Y.; Wang, P.Z.; Wang, H.R.; Li, C.Y. Effects of the state of Co species in Co/Al2O3 catalysts on the catalytic performance of propane dehydrogenation. Appl. Surf. Sci. 2018, 441, 688–693. [Google Scholar] [CrossRef]
  42. Wu, Y.; Long, J.; Wei, S.; Gao, Y.; Yang, D.; Dai, Y.; Yang, Y. Non-oxidative propane dehydrogenation over Co/Ti-ZSM-5 catalysts: Ti species-tuned Co state and surface acidity. Micropor. Mesopor. Mater. 2022, 341, 112115. [Google Scholar] [CrossRef]
  43. Li, Y.; Zhang, Q.; Fu, S.; Kondratenko, V.A.; Otroshchenko, T.; Bartling, S.; Zhang, Y.; Zanina, A.; Wang, Y.; Cui, G.; et al. Active species and fundamentals of their creation in Co-containing catalysts for efficient propane dehydrogenation to propylene. Chem. Eng. J. 2023, 460, 14177. [Google Scholar] [CrossRef]
  44. Jacobson, D.; Freiser, B. Studies of the Reactions of Group 8 Transition-Metal Ions Fe+, Co+, and Ni+ with Linear Alkanes. Determination of Reaction Mechanisms and MCnH2n+ Ion Structures Using Fourier Transform Mass Spectrometry Collision-Induced Dissociation. J. Am. Chem. Soc. 1983, 105, 5197–5206. [Google Scholar] [CrossRef]
  45. Freas, R.B.; Campana, J.E. Dehydrogenation of isobutane by oxygen-deficient cobalt/oxygen cluster ions. J. Am. Chem. Soc. 1986, 108, 4659–4661. [Google Scholar] [CrossRef]
  46. Van Koppen, P.; Brodbelt-Lustig, J.; Bowers, M.; Dearden, D.; Beauchamp, J.; Fisher, E.R.; Armentrout, P. Transition Metal Ion Mediated C–H and C–C Bond Activation of Alkanes: Dynamical Coupling between Entrance and Exit Channel Transition States. J. Am. Chem. Soc. 1991, 113, 2359–2369. [Google Scholar] [CrossRef]
  47. Van Koppen, P.A.; Bowers, M.T.; Haynes, C.L.; Armentrout, P.B. Reactions of Ground-State Ti+ and V+ with Propane: Factors That Govern C–H and C–C Bond Cleavage Product Branching Ratios. J. Am. Chem. Soc. 1998, 120, 5704–5712. [Google Scholar] [CrossRef]
  48. Gao, K.; Yoshikai, N. Low-Valent Cobalt Catalysis: New Opportunities for C−H Functionalization. Acc. Chem. Res. 2014, 47, 1208–1219. [Google Scholar] [CrossRef]
  49. Rodriguez, J.A.; Hanson, J.C.; Frenkel, A.I.; Kim, J.Y.; Perez, M. Experimental and Theoretical Studies on the Reaction of H2 with NiO: Role of O Vacancies and Mechanism for Oxide Reduction. J. Am. Chem. Soc. 2002, 124, 346–354. [Google Scholar] [CrossRef]
  50. Jeangros, J.; Hansen, T.W.; Wagner, J.B.; Damsgaard, C.D.; Dunin-Borkowski, R.E.; Hebert, C.; Van herle, J.; Hessler-Wyser, A. Reduction of nickel oxide particles by hydrogen studied in an environmental TEM. J. Mater. Sci. 2012, 48, 2893–2907. [Google Scholar] [CrossRef]
  51. Rasi, N.M.; Ponnurangam, S.; Mahinpey, N. First-principles investigations into the effect of oxygen vacancies on the enhanced reactivity of NiO via Bader charge and density of states analysis. Catal. Today 2023, 407, 172–181. [Google Scholar] [CrossRef]
  52. Li, Y.; Chen, T.; Zhao, S.; Wu, P.; Chong, Y.; Li, A.; Zhao, Y.; Chen, G.; Jin, X.; Qiu, Y.; et al. Engineering Cobalt Oxide with Coexisting Cobalt Defects and Oxygen Vacancies for Enhanced Catalytic Oxidation of Toluene. ACS Catal. 2022, 12, 4906–4917. [Google Scholar] [CrossRef]
  53. Kumar, J.; Jung, H.J.; Neiber, R.R.; Soomro, R.A.; Kwon, Y.J.; Hassan, N.U.; Shon, M.; Lee, J.H.; Baek, K.-Y.; Cho, K.Y. Recent advances in oxygen deficient metal oxides: Opportunities as supercapacitor electrodes. Int. J. Energy Res. 2022, 46, 7055–7081. [Google Scholar] [CrossRef]
  54. Liu, H.; Fu, H.; Liu, Y.; Chen, X.; Yu, K.; Wang, L. Synthesis, characterization and utilization of oxygen vacancy contained metal oxide semiconductors for energy and environmental catalysis. Chemosphere 2021, 272, 129534. [Google Scholar] [CrossRef] [PubMed]
  55. Zhang, X.; Liu, X.; Zeng, Y.; Tong, Y.; Lu, X. Oxygen Defects in Promoting the Electrochemical Performance of Metal Oxides for Supercapacitors: Recent Advances and Challenges. Small Methods 2020, 4, 1900823. [Google Scholar] [CrossRef]
  56. Zeeshan, M.; Chang, Q.-Y.; Zhang, J.; Hu, P.; Sui, Z.-J.; Zhou, X.-G.; Chen, D.; Zhu, Y.-A. Effects of Oxygen Vacancy and Pt Doping on the Catalytic Performance of CeO2 in Propane Dehydrogenation: A First-Principles Study. Chin. J. Chem. 2021, 39, 2391–2402. [Google Scholar] [CrossRef]
  57. Xie, Z.; Yu, T.; Song, W.; Li, J.; Zhao, Z.; Liu, B.; Gao, Z.; Li, D. Highly active nanosized anatase TiO2–x oxide catalysts in situ formed through reduction and Ostwald ripening processes for propane dehydrogenation. ACS Catal. 2020, 10, 14678–14693. [Google Scholar] [CrossRef]
  58. Li, C.-F.; Guo, X.; Shang, Q.-H.; Yan, X.I.; Ren, C.; Lang, W.-Z.; Guo, Y.-J. Defective TiO2 for propane dehydrogenation. Ind. Eng. Chem. Res. 2020, 59, 4377–4387. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zhao, Y.; Otroshchenko, T.; Lund, H.; Pohl, M.-M.; Rodemerck, U.; Linke, D.; Jiao, H.; Jiang, G.; Kondratenko, E.V. Control of coordinatively unsaturated Zr sites in ZrO2 for efficient C–H bond activation. Nat. Commun. 2018, 9, 3794. [Google Scholar] [CrossRef]
  60. Xie, Z.; Li, Z.; Tang, P.; Song, Y.; Zhao, Z.; Kong, L.; Fan, X.; Xiao, X. The effect of oxygen vacancies on the coordinatively unsaturated Al–O acid-base pairs for propane dehydrogenation. J. Catal. 2021, 397, 172–182. [Google Scholar] [CrossRef]
  61. Yun, Y.; Araujo, J.R.; Melaet, G.; Baek, J.; Archanjo, B.S.; Oh, M.; Alivisatos, A.P.; Somorjai, G.A. Activation of tungsten oxide for propane dehydrogenation and its high catalytic activity and selectivity. Catal. Lett. 2017, 147, 622–632. [Google Scholar] [CrossRef]
  62. Golovnya, R.V.; Samusenko, A.L.; Mistryukov, E.A. Analysis of polar compounds on PEG-40M/KF glass capillary columns. J. High Resolut. Chromatogr. 1979, 2, 609–612. [Google Scholar] [CrossRef]
Figure 1. Powder XRD patterns of as-synthesized and spent–regenerated catalysts.
Figure 1. Powder XRD patterns of as-synthesized and spent–regenerated catalysts.
Catalysts 13 01419 g001
Figure 2. H2-TPR profiles of (1) CoAl-0.5, (2) CoAl-0.25, (3) CoAl-0.25-R, (4) CoAl-0.1, (5) CoAl-0.1-R, (6) CoAl-0.05 and (7) Co/Al2O3 catalysts.
Figure 2. H2-TPR profiles of (1) CoAl-0.5, (2) CoAl-0.25, (3) CoAl-0.25-R, (4) CoAl-0.1, (5) CoAl-0.1-R, (6) CoAl-0.05 and (7) Co/Al2O3 catalysts.
Catalysts 13 01419 g002
Figure 3. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of as-synthesized CoAl-0.1 catalyst.
Figure 3. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of as-synthesized CoAl-0.1 catalyst.
Catalysts 13 01419 g003aCatalysts 13 01419 g003b
Figure 4. Low- (a) and higher- (b) magnification TEM and (c) HAADF-STEM images of as-synthesized CoAl-0.25 catalyst, representing a variety of phases, such as crystalline alumina, cobalt aluminate and cobalt oxides (including small particles on alumina).
Figure 4. Low- (a) and higher- (b) magnification TEM and (c) HAADF-STEM images of as-synthesized CoAl-0.25 catalyst, representing a variety of phases, such as crystalline alumina, cobalt aluminate and cobalt oxides (including small particles on alumina).
Catalysts 13 01419 g004
Figure 5. High-resolution Co 2p XPS spectra of (1) CoAl-0.5, (2) CoAl-0.25, (3) CoAl-0.25-R, (4) CoAl-0.1, (5) CoAl-0.1-R, (6) CoAl-0.05 and (7) Co/Al2O3 catalysts.
Figure 5. High-resolution Co 2p XPS spectra of (1) CoAl-0.5, (2) CoAl-0.25, (3) CoAl-0.25-R, (4) CoAl-0.1, (5) CoAl-0.1-R, (6) CoAl-0.05 and (7) Co/Al2O3 catalysts.
Catalysts 13 01419 g005
Figure 6. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl catalysts at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Figure 6. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl catalysts at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Catalysts 13 01419 g006
Figure 7. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl-0.1, CoAl-0.1-R (spent–regenerated) and CoAl-0.1-H (pre-treated with H2) catalysts at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Figure 7. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl-0.1, CoAl-0.1-R (spent–regenerated) and CoAl-0.1-H (pre-treated with H2) catalysts at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Catalysts 13 01419 g007
Figure 8. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl-0.25 and spent–regenerated CoAl-0.25-RN catalysts (where N is the execution number) at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Figure 8. Propane conversion (a) and selectivity to propylene (b) as functions of time on stream over CoAl-0.25 and spent–regenerated CoAl-0.25-RN catalysts (where N is the execution number) at T = 600 °C and P = 1 atm. Gas mixture: 10 vol.% propane, N2 balance; GHSV = 7500 mL h−1 gcat−1.
Catalysts 13 01419 g008
Figure 9. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of the spent CoAl-0.1 catalyst (after 3 h of operation at T = 600 °C and P = 1 atm for the gas mixture of 10 vol% propane, N2 balance, GHSV = 7500 mL h−1 gcat−1).
Figure 9. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of the spent CoAl-0.1 catalyst (after 3 h of operation at T = 600 °C and P = 1 atm for the gas mixture of 10 vol% propane, N2 balance, GHSV = 7500 mL h−1 gcat−1).
Catalysts 13 01419 g009
Figure 10. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of the spent–regenerated CoAl-0.1-R catalyst.
Figure 10. (a,b) TEM, (c,d) HAADF-STEM and (e,f) related EDX-STEM Co and Al elemental mapping images of the spent–regenerated CoAl-0.1-R catalyst.
Catalysts 13 01419 g010
Scheme 1. Probable mechanism of PDH involving OVs generated on the surface of tetrahedral Co+2 oxide forms.
Scheme 1. Probable mechanism of PDH involving OVs generated on the surface of tetrahedral Co+2 oxide forms.
Catalysts 13 01419 sch001
Table 1. Textural properties of the catalysts.
Table 1. Textural properties of the catalysts.
CatalystCo/Al 1Co (wt%) 2SBET
(m2 g−1)
Pore Size
(nm)
Vtotal
(cm3 g−1)
1CoAl-0.51:228.815212.90.49
2CoAl-0.251:417.021916.50.90
3CoAl-0.25-R 31:413.213721.70.74
4CoAl-0.11:106.729114.81.07
5CoAl-0.1-R 31:106.919119.00.91
6CoAl-0.051:204.327416.31.12
7Co/Al2O3-3.81776.00.26
8Al2O3--1667.30.30
1 The molar ratio of Co(NO3)2 and Al(NO3)3 taken for the catalyst preparation. 2 Determined by ICP. 3 The spent–regenerated (re-oxidized) catalyst.
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

Chernov, A.N.; Cherepanova, S.V.; Gerasimov, E.Y.; Prosvirin, I.P.; Zenkovets, G.A.; Shutilov, A.A.; Gorbunova, A.S.; Koltunov, K.Y.; Sobolev, V.I. Propane Dehydrogenation over Cobalt Aluminates: Evaluation of Potential Catalytic Active Sites. Catalysts 2023, 13, 1419. https://doi.org/10.3390/catal13111419

AMA Style

Chernov AN, Cherepanova SV, Gerasimov EY, Prosvirin IP, Zenkovets GA, Shutilov AA, Gorbunova AS, Koltunov KY, Sobolev VI. Propane Dehydrogenation over Cobalt Aluminates: Evaluation of Potential Catalytic Active Sites. Catalysts. 2023; 13(11):1419. https://doi.org/10.3390/catal13111419

Chicago/Turabian Style

Chernov, Aleksey N., Svetlana V. Cherepanova, Evgeny Yu. Gerasimov, Igor P. Prosvirin, Galina A. Zenkovets, Alexei A. Shutilov, Anna S. Gorbunova, Konstantin Yu. Koltunov, and Vladimir I. Sobolev. 2023. "Propane Dehydrogenation over Cobalt Aluminates: Evaluation of Potential Catalytic Active Sites" Catalysts 13, no. 11: 1419. https://doi.org/10.3390/catal13111419

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