Next Article in Journal
Iron Hydroxide/Oxide-Reduced Graphene Oxide Nanocomposite for Dual-Modality Photodynamic and Photothermal Therapy In Vitro and In Vivo
Next Article in Special Issue
The 10th Anniversary of Nanomaterials—Recent Advances in Environmental Nanoscience and Nanotechnology
Previous Article in Journal
Titanium Dioxide Induces Apoptosis under UVA Irradiation via the Generation of Lysosomal Membrane Permeabilization-Dependent Reactive Oxygen Species in HaCat Cells
Previous Article in Special Issue
Propane Steam Reforming over Catalysts Derived from Noble Metal (Ru, Rh)-Substituted LaNiO3 and La0.8Sr0.2NiO3 Perovskite Precursors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Support Effects on the Activity of Ni Catalysts for the Propane Steam Reforming Reaction

by
Aliki Kokka
1,
Athanasia Petala
2 and
Paraskevi Panagiotopoulou
1,*
1
School of Chemical and Environmental Engineering, Technical University of Crete, 73100 Chania, Greece
2
Department of Chemical Engineering, University of Patras, 26504 Patras, Greece
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(8), 1948; https://doi.org/10.3390/nano11081948
Submission received: 30 June 2021 / Revised: 19 July 2021 / Accepted: 27 July 2021 / Published: 28 July 2021

Abstract

:
The catalytic performance of supported Ni catalysts for the propane steam reforming reaction was investigated with respect to the nature of the support. It was found that Ni is much more active when supported on ZrO2 or YSZ compared to TiO2, whereas Al2O3− and CeO2-supported catalysts exhibit intermediate performance. The turnover frequency (TOF) of C3H8 conversion increases by more than one order of magnitude in the order Ni/TiO2 < Ni/CeO2 < Ni/Al2O3 < Ni/YSZ < Ni/ZrO2, accompanied by a parallel increase of the selectivity toward the intermediate methane produced. In situ FTIR experiments indicate that CHx species produced via the dissociative adsorption of propane are the key reaction intermediates, with their hydrogenation to CH4 and/or conversion to formates and, eventually, to CO, being favored over the most active Ni/ZrO2 catalyst. Long term stability test showed that Ni/ZrO2 exhibits excellent stability for more than 30 h on stream and thus, it can be considered as a suitable catalyst for the production of H2 via propane steam reforming.

Graphical Abstract

1. Introduction

The high energy efficiency of fuel cells has drawn considerable attention toward the development of hydrogen production technologies. Hydrogen can be produced from fossil fuels either via hydrocarbon pyrolysis or hydrocarbon reforming processes including steam reforming, partial oxidation or autothermal reforming [1,2,3,4,5]. Biomass processes consisting of biological (bio-photolysis, dark fermentation, photo fermentation) and thermochemical (pyrolysis, gasification, combustion, liquefaction) methods as well as water splitting processes such as electrolysis, thermolysis or photolysis can be alternatively applied for the production of H2 from renewable energy sources [2,6]. However, the latter approaches are facing some major obstacles mostly related to high cost and low H2 yields. Currently, steam reforming of light hydrocarbons, including natural gas, ethane, propane, butane and liquified petroleum gas (LPG), are considered among the most promising and economical routes for hydrogen production [7]. Propane, which is the main component of LPG, has many advantages such as high energy density, compressibility to a transportable liquid at normal temperature and well-developed infrastructure which enable its use worldwide [8,9,10]. Moreover, propane can be stored and transferred as LPG through a wide distribution network or in high pressure cylinders in order to be supplied in remote places (e.g., agricultural, inaccessible or camping areas) or for domestic uses (e.g., households) [3,7].
Under propane steam reforming conditions, the water-gas shift reaction occurs simultaneously at low temperatures contributing to H2 and CO2 production, whereas CO/CO2 methanation may also run in parallel yielding CH4 and H2O. Methane can be also formed via hydrogenation of CHx species derived by the dissociative adsorption of propane on the catalyst surface or through propane decomposition accompanied by ethylene production. In certain cases, the C2H4, CH4 and CO thus produced are further decomposed leading to the formation of coke on the catalyst surface and consequently, to its progressive deactivation [11,12].
Supported noble metal (Rh, Ru, Pt, Pd) catalysts have been proposed to efficiently catalyze the production of H2 via propane steam reforming exhibiting high resistance to coke formation [3,11,13,14]. However, the high cost and low availability of noble metals are major drawbacks restricting their use in practical applications [15,16,17]. Noble metals can be replaced by nickel, which is less expensive and able to convert propane with high H2 yields. However, Ni is susceptible to coke formation and particles sintering considered to be responsible for catalyst deactivation [5,10,17]. The lifetime of Ni-based catalysts can be improved by optimization of the reaction conditions, catalyst promotion, improvement of the catalyst synthesis method as well as by the proper selection of the support [3,9,10,12,16,18,19].
Regarding the support material, it has been proposed that metal oxides characterized by high oxygen storage capacity, such as CeO2, YSZ, TiO2, ZrO2 or CeO2-ZrO2, are able to suppress carbon deposition through the participation of their lattice oxygen in carbon removal [3,16]. Moreover, Ni catalysts supported on metal oxides or promoted metal oxides, which favor steam adsorption and mobility of surface hydroxyls, have been found to facilitate coke gasification [17,20].
Metal-support interactions have been also reported to impose a dramatic effect on both carbon deposition and metal particles sintering under conditions of hydrocarbons reforming [5,21,22]. It was demonstrated that stronger interactions between Ni and MgAl2O4 support resulted in high Ni dispersion and inhibition of the formation of large Ni clusters [22], whereas weak interactions between Ni and SiO2 carrier were found to accelerate sintering and coke formation [21].
Therefore, the economic viability and practical applicability of H2 production via propane steam reforming may be facilitated by the development of efficient Ni catalysts supported on suitable metal oxides, which will be able to possess both high activity and resistance against carbon deposition and particles sintering to realize economically viable reforming processes. In the present study the effect of the nature of the support (TiO2, CeO2, Al2O3, YSZ, ZrO2) on the activity and selectivity of Ni-based catalysts for the propane steam reforming reaction was investigated. Mechanistic aspects related to the support influence on the reaction pathway were also studied and are discussed.

2. Materials and Methods

2.1. Catalyst Preparation and Characterization

The wet impregnation method was applied to prepare Ni (5 wt.%) catalysts supported on commercial metal oxide powders by using an aqueous solution of Ni (Ni(NO3)2∙6H2O) as the metal precursor salt. The commercial metal oxide carriers were used as received and were (a) activated aluminum oxide (Al2O3), catalyst support, 99% (metals basis) (Alfa Aesar, Kandel, Germany), (b) AEROXIDE® TiO2 P25 (TiO2) (Evonik Industries AG, Essen, Germany), (c) cerium (IV) oxide (CeO2), nanopowder, 99.5% min (Alfa Aesar, Kandel, Germany), (d) yttria-stabilized zirconia (YSZ) (8Y-SZ, Tosoh, Amsterdam, The Netherlands) and (e) zirconium (IV) oxide (ZrO2) 99% (metal basis) (Alfa Aesar, Kandel, Germany). The resulting materials were dried at 110 °C for 24 h followed by reduction at 400 °C under H2 flow for 2 h. The selection of reducing Ni catalysts at 400 °C for 2 h was based on previous studies which indicated that under these reducing conditions nickel oxide species are able to be completely converted to metallic nickel [23,24,25].
The X-ray diffraction patterns of the catalysts were recorded using an X-ray powder diffractometer (A Brucker D8 Advance instrument, Bruker, Karlsruhe, Germany) using Cu Ka radiation (λ = 0.15406 nm, 40 kV, 40 mA) and a scan rate of 0.025°/s over a range of 2θ between 20 and 80°. The diffraction pattern was identified by comparison with those supplied from the JCPDS data base, whereas the primary crystallite size of MxOy (dMxOy) was estimated according to Scherrer’s equation:
d M x O y = 0.9 λ Β c o s θ
where θ is the angle of diffraction corresponding to the peak broadening, B is the full-width at half maximum intensity (in radians) and λ = 0.15406 nm is the X-ray wavelength corresponding to CuKa radiation.
The specific surface area (SSA) of the supported Ni catalysts were measured by N2 adsorption at 77 K (B.E.T. technique) using a Gemini III 2375 instrument (Micromeritics, Norcross, GA, USA). Carbon monoxide chemisorption measurements at 25 °C were applied for the determination of Ni dispersion and mean particle size using a modified Sorptomatic 1900 apparatus (Fisons Instruments, Glaskow, UK) and assuming a CO:Me stoichiometry of 1:1, an atomic surface area of 6.5 Å2 and spherical particles. CO chemisorption measurements were used instead of H2 chemisorption in order to avoid overestimation of Ni dispersion due to hydrogen spillover effects, which have been previously found to occur over supported Ni catalysts [26,27]. Nickel particle size was calculated according to the following equation:
d N i = 60000 ρ N i S N i   [ Å ]
where dNi is the mean crystallite diameter, ρNi (= 8.9 g∙cm−3) is the density of Ni and SNi [m2/gNi] is the surface area per gram of Ni.
Transmission electron microscopy (TEM) images were obtained with a JEM-2100 system (JEOL, Akishima, Tokyo, Japan) operated at 200 kV (point resolution 0.23 nm) using an Erlangshen CCD Camera (Model 782 ES500W, Gatan Inc., Pleasanton, CA, USA). Samples were dispersed in water and spread onto a carbon-coated copper grid (200 mesh). Details related to the equipment and procedures used for catalyst characterization have been described in detail elsewhere [28].

2.2. Catalytic Performance Tests and Kinetic Measurements

The catalytic performance of the synthesized materials was studied in a tubular fixed-bed quartz reactor under atmospheric pressure using an apparatus which has been described in detail elsewhere [11]. The reaction conditions were as follows: temperature range 400–750 °C, H2O/C = 3.25, and gas hourly space velocity (GHSV) = 55,900 h−1. The reactor was loaded with 150 mg of catalyst (particle diameter: 0.15 < dp < 0.25 mm) and placed in an electric furnace, where it was reduced in situ at 300 °C for 1 h under 50%H2/He flow (60 cm3 min−1) to ensure that the Ni exists in its metallic phase prior to catalytic performance tests. Catalyst reduction was followed by heating at 750 °C under He and subsequent switch of the flow to the feed stream consisted of 4.5%C3H8 + 0.15%Ar + 44%H2O (He balance). Argon was used as internal standard in order to account for the volume change. Water was fed through an HPLC pump (LD Class Pump, TELEDYNE SSI, PA, USA) into a vaporizer maintained at 180 °C and mixed with the gas stream coming from mass-flow controllers. A condenser immersed in an ice bath was placed at the exit of the reactor to condensate water prior to introduction of the gas stream to the analysis system. Reaction gases (He, 30%C3H8−1%Ar/He, H2) are supplied from high-pressure gas cylinders (Buse Gas, Bad Hönningen, Germany) and are of ultrahigh purity. Measurements of reactants’ and products’ concentrations were obtained by stepwise decreasing temperature up to 400 °C. The effluent from the reactor was analyzed using two gas chromatographs (Shimadzu, Kyoto, Japan) which were connected in parallel. The procedure used for gas phase analysis was described in our previous study [11]. The conversion of propane ( X C 3 H 8 ) was calculated using the following expression:
X C 3 H 8 = [ C a r b o n ] t o t a l , o u t [ C a r b o n ] t o t a l , o u t + [ C 3 H 8 ] o u t × 100
where [Carbon]total,out is the sum of the concentrations of all carbon containing products:
[ C a r b o n ] t o t a l , o u t = [ C O ] + [ C O 2 ] + [ C H 4 ] 3 + 2 × [ C 2 H 4 ] + [ C 2 H 6 ] 3
Selectivity toward reaction products containing carbon was defined using Equation (5). The factor n corresponds to the number of carbon atoms in the corresponding molecule (e.g., for CO is 1, for C2H4 is 2 etc.):
S C n = [ C n ] × n 3 × [ C a r b o n ] t o t a l , o u t × 100
Selectivity toward hydrogen production was defined as the concentration of hydrogen produced divided with the concentration of all products containing hydrogen according to Equation (6). The factor m represents the number of hydrogen atoms in the corresponding molecule (e.g., for CH4 and C2H4 is 4).
S H 2 ( % ) = [ H 2 ] [ H 2 ] + m / 2 × [ C H n m ] × 100
The intrinsic reaction rates for propane steam reforming reaction were measured for low propane conversions ( X C 3 H 8 < 10%) by varying W/F using the following expression:
R C 3 H 8 = [ C 3 H 8 ] i n F i n [ C 3 H 8 ] o u t F o u t W × 100
where RC3H8 is the molar rate of C3H8 consumption (mol s−1 gcat−1), [C3H8]in, [C3H8]out, are the inlet and outlet concentrations (v/v) of C3H8, respectively, Fin and Fout are the total flow rates in the inlet and outlet of the reactor (mols−1), respectively, and W is the mass of catalyst (gcat).
Turnover frequencies (TOFs) of propane conversion were estimated following Equation (8) taking into account the measurements of both the reaction rates and nickel dispersions:
T O F = R C 8 H 8 A W N i D N i X N i
where AWNi is the atomic weight of nickel (gNi/molNi), XNi is the nickel loading (gNi/gcat) and DNi is the dispersion of nickel.

2.3. In Situ FTIR Spectroscopy

In situ Fourier transform infrared (FTIR) experiments were carried out using an iS20 FTIR spectrometer (Nicolet, Thermo Fischer Scientific, Waltham, MA, USA) equipped with an MCT detector, a KBr beam splitter and a diffuse reflectance (DRIFT) sampling system (Specac, Orpington, UK) accompanied by an environmental chamber suitable for the study of diffusely reflecting solid samples in a controlled atmosphere. A flow system equipped with mass flow controllers, a steam saturator and a set of valves used for controlling the gas stream interacted with the catalyst surface, was directly connected to the gas inlet of the environmental chamber.
In a typical experiment, the catalyst powder was placed in the sampling system and heated at 500 °C in flowing helium for 10 min and then reduced under hydrogen flow at 300 °C for 30 min. The flow was then switched to He and the temperature was increased at 500 °C. After remaining 10 min at this temperature the sample was cooled at 100 °C. While cooling, the background spectra were recorded at the desired temperatures. Finally, the flow was switched to the reaction mixture, which consisted of 0.5%C3H8 +5%H2O (in He). Steam was introduced to the system via an independent He line passing through a saturator containing water maintained at 60 °C. The resulting gas mixture was fed to the DRIFT cell through stainless steel tubing maintained at 60 °C by means of heating tapes. A spectrum was collected at 100 °C after 15 min-on-stream followed by a stepwise increase of temperature up to 500 °C. During heating, spectra were recorded at selected temperatures after an equilibration for 15 min. In all experiments, the total flow through the DRIFT cell was 30 cm3 min−1. Reaction gases (He, 2%C3H8/He, H2) are supplied from high-pressure gas cylinders (Buse Gas, Bad Hönningen, Germany) and are of ultrahigh purity.

3. Results

3.1. Catalyst Characterization

The XRD patterns of Ni/MxOy catalysts are shown in Figure 1. The characteristic peaks located at the diffraction angles of 32.7°, 37.7°, 39.9°, 45.8° and 67.5° were appeared for Ni/Al2O3 which are attributed to (220), (311), (222), (400) and (440) facets of cubic Al2O3 (JCPDS Card No. 10-425), respectively. The XRD spectra recorded for Ni/CeO2 catalyst consisted of peaks located at 2θ = 28.79°, 33.26°, 47.62°, 56.34°, 59.1°, 69.42°, 76.66°, 79.11° attributed to (111), (200), (220), (311), (222), (400), (331) and (420) planes of the cubic CeO2 (JCPDS Card No. 2-1306), whereas in the case of Ni/ZrO2 catalyst numerous peaks were detected on the XRD pattern. In particular, the peaks detected at 24.2°, 24.5°, 28.5°, 31.7°, 34.5°, 35.5°, 38.9°, 41.0°, 41.5°, 45.1°, 45.8°, 49.5°, 50.5°, 54.4°, 55.7°, 57.5°, 60.2°, 62.1°, 63.2°, 66.0°, 71.6°, 75.3° correspond to (011), (−110), (−111), (111), (002), (200), (021), (−211), (−121), (112), (211), (022), (−221), (202), (013), (212), (−302), (113), (311), (−321), (−104), (−140) planes of monoclinic ZrO2 (JCPDS Card No. 13-307). When Ni was supported on YSZ, the XRD pattern was characterized by reflections at 30.4°, 35.1°, 50.4°, 59.9°, 62.9° and 74.0° attributed to (111), (200), (220), (311), (222) and (400) planes of YSZ (JCPDS Card No. 82-1246), respectively.
Results obtained from Ni/TiO2 catalyst showed that the sample consisted of TiO2 in both its anatase and rutile form exhibiting peaks at 2θ = 25.6° (101), 37.2° (103), 38.2° (004), 38.9° (112), 48.4° (200), 54.3° (105), 55.4° (211), 63.1° (204), 69.3° (116), 70.7° (220), 75.4° (215) and 76.4° (301) for anatase (JCPDS Card No. 21-1272), and at 2θ = 27.6° (110), 36.3° (101), 41.6° (111), 44.3° (210), 54.6° (211), 56.9° (220) and 64.3° (310) for rutile (JCPDS Card No. 21-1276).
In the case of Ni/TiO2 and Ni/YSZ catalysts an additional weak peak located 44.5° was appeared corresponding to Ni (111) plane (JCPDS Card No. 04-0850). The absence of peaks corresponding to metallic Ni for the rest catalysts investigated is due to the low Ni loading and/or particle size. The primary crystallite size of the supports was estimated according to Scherrer’s formula at the diffraction angles corresponding to (440) plane for Al2O3, (−111) plane for ZrO2, (111) plane for CeO2, (111) plane for YSZ and (101) plane for TiO2, and it was found to be 6.0 nm for Ni/Al2O3, 10.5 nm for Ni/CeO2, 15.0 nm for Ni/ZrO2, 20.9 nm for Ni/YSZ and 21.8 nm for Ni/TiO2 (Table 1).
The SSAs of Ni catalysts supported on metal oxide (MxOy) carriers were estimated equal to 39 m2/g for Ni/ZrO2, 11 m2/g for Ni/YSZ, 66 m2/g for Ni/Al2O3, 39 m2/g for Ni/CeO2 and 41 m2/g for Ni/TiO2 (Table 1).
Results of Ni dispersion (DNi) and mean particle size (dNi) estimated from CO chemisorption meaurements are summarized in Table 1. Generally, low Ni dispersions were estimated for all the investigated catalysts, most possibly due to the high Ni content (5 wt.%) in agreement with previous studies [4,6]. Higher Ni dispersion of 11.9% and smaller particle size of 8.5 nm was found for Ni/CeO2 catalyst, whereas Ni/TiO2 exhibited the lowest value of Ni dispersion of 2.8% and the largest particle size of 36.1 nm.
Figure 2 shows representative TEM images and selected area electron diffraction (SAED) patterns obtained from Ni/YSZ, Ni/CeO2 and Ni/TiO2 catalysts. In all cases Ni particles appear as fairly homogeneously distributed spherical particles with average sizes of 20 nm for Ni/YSZ, 10 nm for Ni/CeO2 and 30 nm for Ni/TiO2, in agreement with those estimated according to CO chemisorption measurements (Table 1).
It should be noted that, based on the results of Table 1, the mean particle size of Ni is similar to the average size of the corresponding metal oxide used as support for all the investigated catalysts. This may hinder distinguishing between Ni particles and the MxOy carrier in TEM images. Thus, SAED analysis was performed to calculate the d-spacing in an attempt to further discerned Ni particles from those of metal oxide support (Figure 2, Table 2). Results indicated that, in all cases, Ni particles are present in TEM images as evidenced by the appearance of the (111) plane of Ni (dspacing = 0.21 nm, JCPDS No 1-1258). The appearance of the (101) (dspacing = 0.35 nm, JCPDS No 1-562) and (103) (dspacing = 0.24 nm, JCPDS No 1-562) planes of TiO2, the (111) (dspacing = 0.30 nm, JCPDS No 30-1468) and (200) (dspacing = 0.26 nm, JCPDS No 30-1468) planes of YSZ, as well as the (111) (dspacing = 0.31 nm, JCPDS No 1-800) and (200) (dspacing = 0.27 nm, JCPDS No 1-800) planes of CeO2 also confirmed the presence of the metal oxides.

3.2. Influence of the Nature of the Support on Catalytic Activity

The influence of the nature of the support on catalytic performance for the propane steam reforming reaction has been investigated using Ni catalysts (5 wt.%) supported on five different commercial metal oxide powders (ZrO2, YSZ, TiO2, Al2O3, CeO2). The results obtained are shown in Figure 3A, where propane conversion is plotted as a function of reaction temperature. It is observed that, among the investigated catalysts, Ni/ZrO2 is the most active one, exhibiting measurable C3H8 conversions at temperatures higher than 400 °C and achieving complete conversion at 750 °C. Although Ni/YSZ is activated at similar temperatures as Ni/ZrO2, the conversion curve of propane is shifted toward higher temperatures. This is also the case for Ni/Al2O3 and Ni/CeO2 catalysts, which present similar performance. The latter catalysts are less active than Ni/YSZ below 550 °C, but are able to reach higher XC3H8 at higher temperatures. The titania-supported catalyst becomes active above 500 °C, with the propane conversion curve being shifted at remarkably higher temperatures. In all examined cases, the carbon balance was satisfactory, with a deviation of <1%.
Results of specific reaction rate measurements are presented in the Arrhenius diagram of Figure 3B, where it is observed that the TOF of propane conversion increases in the order Ni/TiO2 < Ni/CeO2 < Ni/Al2O3 < Ni/YSZ < Ni/ZrO2, with its value at 450 °C being more than one order of magnitude higher when Ni is dispersed on ZrO2 compared to TiO2, and approximately 2.5 times higher than that of Ni/Al2O3. It should be mentioned that, as discussed above, the mean particle size of Ni varies significantly for the investigated catalysts from 8.5 nm for Ni/CeO2 to 36.1 nm for Ni/TiO2. If the Ni particle size were similar for this set of catalysts then the order of catalytic activity could be somewhat different. Interestingly, no trend was observed between the specific reaction rate and Ni particle size or MxOy crystallite size or MxOy surface area. This indicates that either these parameters do not affect catalytic activity or most possibly each of them contributes in a different manner to the reaction rate, resulting in the observed catalyst ranking. It should be noted that all catalysts have been reduced at 400 °C prior to physicochemical characterization measurements. Although the values of SSA or dMxOy or dNi may were different if catalyst pre-reduction was carried out at 750 °C, which is the onset reaction temperature for catalytic performance experiments, the trend of catalytic properties with respect to the nature of the support is not expected to vary due to the catalyst pretreatment at different temperatures, at least to such an extent that would affect the catalyst ranking for the propane steam reforming reaction.
The apparent activation energies (Ea) of the propane steam reforming reaction were calculated from the slopes of the fitted lines of Figure 3B. The results showed that the nature of the metal oxide carrier significantly affects Ea, which takes values between 102 kJ/mol for Ni/CeO2 and 154 kJ/mol for Ni/ZrO2 without presenting any trend with respect to catalytic activity (Table 1). This can be explained taking into account that, as it will be discussed below, several reactions run in parallel under the present experimental conditions each one of which is influenced by the nature of the support in a different manner resulting in the observed random variation of Ea with catalytic activity. The results are in agreement with our previous study where it was found that the apparent activation energy for the reaction of steam reforming of propane over Rh catalysts supported on a variety of metal oxides does not present any trend with the activity order [11].
Figure 4 shows the selectivities toward reaction products as a function of temperature over the supported Ni catalysts investigated. In all cases the main products detected were H2, CO2, CO and CH4 with their selectivities being significantly varied with temperature. In particular, for Ni/ZrO2 catalyst (Figure 4A), both hydrogen (SH2) and CO2 (SCO2) selectivities decrease from 99 to 78% and from 98 to 58%, respectively, with increasing temperature in the range of 390–505 °C followed by an increase of methane selectivity (SCH4) up to 32.5%, indicating the occurrence of CO2 methanation reactions. Carbon dioxide consumption continues with further increase of temperature above 505 °C contrary to SH2 which progressively increases reaching 99% at 720 °C. Consumption of CO2 is followed by production of CO providing evidence that the reverse WGS (RWGS) reaction is enhanced at high temperatures. Moreover, SCH4 decreases above 505 °C and becomes practically zero at 720 °C, implying that the reaction of methane steam reforming occurs contributing to the observed increase of both SH2 and SCO. It should be noted that selectivity toward reaction products containing carbon was defined as the concentration of each product containing carbon at reactor effluent over the concentration of all products containing carbon (5), whereas SH2 was defined as the concentration of hydrogen produced divided by the concentration of all products containing hydrogen (6). Therefore, the values of Scn and SH2 cannot be correlated based on the stoichiometry of the reactions taking place under propane steam reforming conditions.
Qualitatively similar results were obtained for the rest of the investigated Ni catalysts, with the main differences being related to the values of the selectivities toward the reaction products, which reflect the extent of each reaction taking place with respect to the nature of the support. In particular, the observed decrease of SH2 and SCO2 at low temperatures and the simultaneous increase of SCH4 are higher for the most active Ni/ZrO2 (Figure 4A) and Ni/YSZ (Figure 4B) catalysts, followed by Ni/Al2O3 (Figure 4C) and Ni/CeO2 (Figure 4D), whereas it is eliminated for Ni/TiO2 (Figure 4E). As a result methane production increases in the order of Ni/TiO2 < Ni/CeO2 < Ni/Al2O3 < Ni/YSZ < Ni/ZrO2 which is consistent with the order of catalytic activity. This can be clearly seen in Figure 5 where the TOF at 450 °C is plotted as a function of methane selectivity obtained at the same temperature for all the investigated catalysts. It is observed that the specific reaction rate increases from 0.018 s−1 to 0.33 s−1 following the above catalyst ranking and accompanied by a parallel increase of SCH4 from 0 to 29%. The results indicate that there is a clear relationship between catalytic activity and methane production.
It should be noted that besides CO2 hydrogenation, CH4 can be also produced via CO hydrogenation. However, the contribution of the latter reaction does not seem to be significant for the results of the present study taking into account the low SCO below 500 °C and its progressive increase with temperature. Moreover, methane formation may also take place via hydrogenation of CHx species formed following the dissociative adsorption of propane on Ni surface and the subsequent hydrogenation of the so-formed C3Hx species [29,30]. As it will be discussed below, the formation of CHx species intermediates may be the key reaction since it has been proposed that they interact with the hydroxyl groups or lattice oxygen of the support producing CO or CO2 and H2 [3,29,30].
The mass of H2 produced per propane mass unit contained in the feed was calculated at 550 °C and it was found to be higher for Ni/ZrO2 (23.2 wt.%) followed by Ni/CeO2 (21.2 wt.%), Ni/Al2O3 (20.6 wt.%) and Ni/YSZ (19.7 wt.%), whereas Ni/TiO2 exhibits the lowest H2 production (5.2 wt.%).
The influence of the nature of the support on the activity of Ni catalysts for propane steam reforming reaction was also investigated by Harshini et al. [16], who found that Ni/LaAlO3 was more active than Ni/Al2O3, while Ni/CeO2 exhibited intermediate performance. The optimum activity of the former catalyst was attributed to the small Ni nanoparticles dispersed on LaAlO3 surface. Although the effect of the support nature on propane steam reforming activity has not been widely studied over Ni catalysts, certain properties of metal oxide carriers may help explain the results of Figure 3. For example, the use of YSZ as support, which exhibited high activity in the results of the present study, has been found to suppress carbon deposition over Rh-Ni catalysts by providing lattice oxygen, which facilitates carbon removal and enhances the dissociation of C-C bond under reaction conditions [3]. The prevention of coke formation, occurring either via hydrocarbons decomposition or CO dissociation, by the lattice oxygen of the support has been also demonstrated over Ni/CeO2-Al2O3 [19]. Moreover, the addition of manganese oxide on Ni/Al2O3 was found to act as an oxygen donor that is transferred to Ni particles leading to rapid decomposition and oxidation of C3H8 and CH4 or C2H4 that may be produced under reaction conditions, resulting in further H2 production and improvement of the catalyst lifetime [9]. It has been also found that activation of steam followed by H2 formation may be favored over metal catalysts supported on “reducible” metal oxides through generation of oxygen defects, resulting in improved propane steam reforming activity and resistance to coke formation [3,13,31]. Based on previous studies, the reducibility of the supports used in the present study is expected to vary significantly. It is well known that Al2O3 is a hardly reducible metal oxide characterized by low oxygen storage capacity contrary to TiO2 and CeO2, which are strongly reducible metal oxides, or ZrO2 and YSZ, which are characterized by intermediate oxygen mobility [32]. Based on the above, Ni/TiO2 should be also active for the title reaction, taking into account that titania support is characterized by high oxygen storage capacity [32,33]. However, the results of Figure 3 clearly show that: (a) this catalyst was the least active one and (b) the catalytic activity is increased in the order Ni/TiO2 < Ni/CeO2 < Ni/Al2O3 < Ni/YSZ < Ni/ZrO2, which cannot be correlated with the reducibility of the support. Therefore, it is evident that support reducibility is not among the key parameters affecting the catalytic activity of Ni according to the results of the present study. The low activity of Ni/TiO2 catalyst may be related to the fact that Ni/TiO2 has lower Ni dispersion and larger Ni particles, which was previously suggested to suppress both propane steam reforming [15,16], and the intermediate (CO2 or CHx) hydrogenation reactions, in excellent agreement with the results of our previous study [28]. However, large Ni particles may not be solely responsible for the low activity of Ni/TiO2 taking into account that no trend was observed between TOF and Ni particle size for the investigated catalysts.
It should be mentioned that a different activity order was reported in our previous study over supported Rh catalysts, where it was found that Rh/TiO2 was the most active catalyst with TOF being one order of magnitude higher compared to that measured for Rh/CeO2 [11]. This implies that the nature of the metallic phase may affect metal/support interactions leading to variations on propane steam reforming activity and/or possibly changes on the type of active sites on the catalyst surface.

3.3. Long Term Stability Test

The long-term stability of Ni/ZrO2 catalyst, which exhibited the highest activity, was investigated at 650 °C using the same experimental conditions as those used in catalytic performance tests. In this experiment, the catalyst was reduced in situ at 300 °C under 50%H2/He flow followed by heating at 650 °C. The flow was then switched to the reaction mixture and determination of the conversion of propane and product selectivity started. The system was shut down overnight, while the catalyst was kept at room temperature under a He flow. The next day the catalyst is heated to 650 °C in the He flow, followed by switching of the flow to the reaction mixture and determination of XC3H8 and the product selectivity as a function of time. Results obtained are shown in Figure 6, where XC3H8 and SH2, SCO2, SCO and SCH4 are plotted as functions of time-on-stream. As it can be seen the catalyst presents excellent stability for more than 30 h-on-stream. Propane conversion and hydrogen selectivity were varied in the range of 95–99% and 97–98%, respectively. The selectivity toward methane was low (3–4%) whereas SCO and SCO2 exhibited similar values ranging between 46 and 50%. The carbon balance was found to be satisfactory during the stability test, with a deviation lower than 1–2%.

3.4. DRIFT Studies

The interaction of selected catalysts with the reaction mixture was also investigated employing in situ FTIR spectroscopy. Experiments were conducted in the temperature range of 100–500 °C using a feed composition of 0.5%C3H8 + 5%H2O (in He) and the results obtained are shown in Figure 7. It is observed that the spectrum recorded at 100 °C (Figure 7A, trace a) for the pre-reduced Ni/TiO2 catalyst is characterized by two negative bands at 3787 and 3676 cm−1 which can be attributed to losses of ν (OH) intensity of at least two different types of free hydroxyl groups, which are either originally present on TiO2 surface or created via H2O adsorption. Two weak peaks were also detected in the ν (C-H) region, located at 2987 and 2966 cm−1 (trace a) due to C-H stretching vibrations in methyl groups (CH3,ad) and to symmetric C-H vibrations in methylene groups (CH2,ad), respectively [20,31,34,35,36]. These peaks are more obvious in Figure 8A (trace a) where selected spectra in the narrow range of 3200–2400 cm−1 are presented. Moreover, a band at 1642 cm−1 followed by a shoulder at 1560 cm−1 can be discerned, which have been previously assigned to carbonate species associated with TiO2 support [37,38,39,40,41,42]. An increase of temperature results in progressive separation of the latter two bands which are both shifted toward lower wavenumbers. A new band at 1430 cm−1 can be also discerned in the spectra obtained at 350 °C (trace f) which is also due to carbonate species. This peak may be also present in the spectra obtained at lower temperatures but couldn’t be clearly observed due to the low signal-to-noise ratio in the region below 1700 cm−1. The intensities of bands assigned to carbonate species are progressively decreased above 200 °C. This decrease is accompanied by the detection of a weak peak at 2021 cm−1 [38,43,44,45,46], which is characteristic of linear-bonded CO on reduced nickel sites (Ni°), indicating that carabonate species are further decomposed yielding CO and most possibly also CO2 in the gas phase. The weak bands in the ν (C-H) region are present on the spectra obtained up to 500 °C (Figure 7A, trace i) implying that CHx species are thermally stable and remained adsorbed on the catalyst surface.
A similar experiment was conducted over the most active Ni/ZrO2 catalyst and the results obtained are presented in Figure 7B. It is observed that the interaction of catalyst with the reformate mixture at 100 °C (trace a) results in the appearance of bands corresponding to bicarbonate species (1640 cm−1) [47,48,49], CHx species (2983 and 2966 cm−1) [31,36,47] as well as by negative bands (3749 and 3675 cm−1) related to the consumption of surface OH groups [31,36,48]. Increase of temperature at 200 °C (trace c) leads to the progressive development of two bands at 1540 and 1425 cm−1 due to bicarbonate species [47,48,49]. The intensity of the latter bands increases with increasing temperature up to 300 °C and diminishes upon further heating at 350–400 °C. This is also the case for the band at 1640 cm−1, indicating that bicarbonate species are decomposed above 450 °C. At temperatures higher than 400 °C (traces h-i) two new broad bands seem to be developed at ca 1540 and 1350 cm−1. The broadness of these bands implies that they may contain contributions from more than one species with their corresponding bands being overlapped.
This can be clearly seen in the spectrum obtained at 500 °C (trace i) where four bands can be clearly discerned located at 1558, 1520, 1370 and 1331 cm−1. Those detected at 1520 and 1331 cm−1 have been previously attributed to bidentate carbonates [49,50], whereas those located at 1558 and 1370 cm−1 can be assigned to bidentate formate species [51] adsorbed on ZrO2 surface. The appearance of the latter bands is accompanied by evolution of three peaks in the ν(CO) region due to CO linearly adsorbed on reduced Ni sites (2021 cm−1) and bridged bonded CO (1909 and 1858 cm−1) [43,46,52,53,54].
Interestingly, CHx species are eliminated from the spectra obtained above 400 °C followed by evolution of CH4 in the gas phase, as evidenced by the detection of the 3016 cm−1 band (traces h–i). Production of CH4 at the expense of CHx species can be clearly seen in Figure 8B where the spectra obtained at 100 and 450 °C in the wavenumber range of 3200–2400 cm−1 are presented.
Based on the above it can be suggested that the reaction of steam reforming of propane over Ni/ZrO2 catalyst proceeds via a dissociative adsorption of propane on metallic Ni leading to the formation of C3Hx species, which are further decomposed toward CHx species and probably carbon oxides due to the presence of H2O adsorbed on the support surface. This may result in the formation of the bicarbonate species (1640, 1540 and 1425 cm−1) detected at low temperatures on the surface of the support [20]. Part of CHx species are hydrogenated above 400 °C, yielding methane in the gas phase (band at 3016 cm−1) whereas the rest interact with adsorbed water producing formates (bands at 1558 and 1370 cm−1) and, eventually, CO species adsorbed on the Ni surface (bands at 2021, 1909 and 1858 cm−1). Formates may also interact with hydroxyl groups producing H2 and carbonate species (1520 and 1331 cm−1), which are further decomposed to CO2 [20]. It should be noted, however, that, under certain conditions, CHx species may be also dehydrogenated producing C and H2. Surface carbon is either accumulated on the catalyst surface resulting in catalyst deactivation (which is not the case here) or it interacts with the hydroxyl groups or the lattice oxygen of the support yielding CO or CO2 [3,16,19,31,55].
The ability of CHx species to be converted to methane and/or formates on the surface of Ni/ZrO2 may imply that CHx species are more weakly adsorbed and/or more reactive on the surface of this catalyst, thus resulting in higher overall activity for the propane steam reforming reaction. This agrees well with results of Figure 5, where it was shown that catalytic activity increases progressively with increasing methane selectivity. If as discussed above the origin of adsorbed CO is formate species, the high reactivity of CHx species over Ni/ZrO2 is also indirectly confirmed by the significantly higher population of carbonyl species observed over this catalyst. On the other hand, although CHx species were also detected on the surface of Ni/TiO2 catalyst, no band due to gas phase methane or formate species was observed, indicating that CHx species cannot be further converted in the temperature range investigated. The increase of SCH4 in parallel with the increase of catalytic activity (Figure 5) indicates that the reactivity of CHx species is strongly related to the conversion of propane to the desired products. Although detailed mechanistic studies should be conducted to further explore the reaction pathway, it can be suggested that CHx species are key reaction intermediates for the reaction of propane steam reforming.

4. Conclusions

The results of the present study show that the catalytic activity of Ni and the product distribution for the propane steam reforming reaction depend strongly on the nature of the support. The specific reaction rate measured for Ni/ZrO2 catalyst was found to be more than one order of magnitude higher compared to that measured for Ni/TiO2. The intermediate production of CH4 is strongly influenced by the type of metal oxide used as support following the same trend with that of catalytic activity. DRIFTS studies provided evidences that the most active Ni/ZrO2 catalyst is able to convert the intermediate produced CHx species to CH4 and/or formate species and, subsequently, to carbon oxides and H2. In contrast, CHx species seem to be less reactive when Ni is dispersed on TiO2, thus resulting in a lower reaction rate. In addition to the high activity of Ni/ZrO2 catalyst, it was also found to be stable for more than 30 h on stream and therefore, is a promising candidate for the production of H2 for fuel cell applications.

Author Contributions

Conceptualization, P.P.; methodology, P.P.; investigation, A.K., A.P. and P.P.; data curation, A.K., A.P. and P.P.; writing—original draft preparation, P.P.; writing—review and editing, P.P.; visualization, P.P.; supervision, P.P.; project administration, P.P.; funding acquisition, P.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH–CREATE–INNOVATE (project code: T1EDK–02442).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Kalamaras, C.M.; Efstathiou, A.M. Hydrogen Production Technologies: Current State and Future Developments. Conf. Pap. Energy 2013, 2013, 690627. [Google Scholar] [CrossRef] [Green Version]
  2. Nikolaidis, P.; Poullikkas, A. A comparative overview of hydrogen production processes. Renew. Sustain. Energy Rev. 2017, 67, 597–611. [Google Scholar] [CrossRef]
  3. Im, Y.; Lee, J.H.; Kwak, B.S.; Do, J.Y.; Kang, M. Effective hydrogen production from propane steam reforming using M/NiO/YSZ catalysts (M = Ru, Rh, Pd, and Ag). Catal. Today 2018, 303, 168–176. [Google Scholar] [CrossRef]
  4. Santamaria, L.; Lopez, G.; Arregi, A.; Amutio, M.; Artetxe, M.; Bilbao, J.; Olazar, M. Influence of the support on Ni catalysts performance in the in-line steam reforming of biomass fast pyrolysis derived volatiles. Appl. Catal. B Environ. 2018, 229, 105–113. [Google Scholar] [CrossRef]
  5. Santamaria, L.; Lopez, G.; Arregi, A.; Amutio, M.; Artetxe, M.; Bilbao, J.; Olazar, M. Stability of different Ni supported catalysts in the in-line steam reforming of biomass fast pyrolysis volatiles. Appl. Catal. B Environ. 2019, 242, 109–120. [Google Scholar] [CrossRef]
  6. Miyazawa, T.; Kimura, T.; Nishikawa, J.; Kado, S.; Kunimori, K.; Tomishige, K. Catalytic performance of supported Ni catalysts in partial oxidation and steam reforming of tar derived from the pyrolysis of wood biomass. Catal. Today 2006, 115, 254–262. [Google Scholar] [CrossRef]
  7. Al-Zuhair, S.; Hassan, M.; Djama, M.; Khaleel, A. Hydrogen Production by Steam Reforming of Commercially Available LPG in UAE. Chem. Eng. Commun. 2017, 204, 141–148. [Google Scholar] [CrossRef]
  8. Do, J.Y.; Lee, J.H.; Park, N.-K.; Lee, T.J.; Lee, S.T.; Kang, M. Synthesis and characterization of Ni2−xPdxMnO4/γ-Al2O3 catalysts for hydrogen production via propane steam reforming. Chem. Eng. J. 2018, 334, 1668–1678. [Google Scholar] [CrossRef]
  9. Do, J.Y.; Park, N.-K.; Lee, T.J.; Lee, S.T.; Kang, M. Effective hydrogen productions from propane steam reforming over spinel-structured metal-manganese oxide redox couple catalysts. Int. J. Energy Res. 2018, 42, 429–446. [Google Scholar] [CrossRef]
  10. Barzegari, F.; Kazemeini, M.; Farhadi, F.; Rezaei, M.; Keshavarz, A. Preparation of mesoporous nanostructure NiO–MgO–SiO2 catalysts for syngas production via propane steam reforming. Int. J. Hydrogen Energy 2020, 45, 6604–6620. [Google Scholar] [CrossRef]
  11. Kokka, A.; Katsoni, A.; Yentekakis, I.V.; Panagiotopoulou, P. Hydrogen production via steam reforming of propane over supported metal catalysts. Int. J. Hydrogen Energy 2020. [Google Scholar] [CrossRef]
  12. Kokka, A.; Ramantani, T.; Panagiotopoulou, P. Effect of Operating Conditions on the Performance of Rh/TiO2 Catalyst for the Reaction of LPG Steam Reforming. Catalysts 2021, 11, 374. [Google Scholar] [CrossRef]
  13. Yu, L.; Sato, K.; Nagaoka, K. Rh/Ce0.25Zr0.75O2 Catalyst for Steam Reforming of Propane at Low Temperature. ChemCatChem 2019, 11, 1472–1479. [Google Scholar] [CrossRef]
  14. Karakaya, C.; Karadeniz, H.; Maier, L.; Deutschmann, O. Surface Reaction Kinetics of the Oxidation and Reforming of Propane over Rh/Al2O3 Catalysts. ChemCatChem 2017, 9, 685–695. [Google Scholar] [CrossRef]
  15. Aghamiri, A.R.; Alavi, S.M.; Bazyari, A.; Azizzadeh Fard, A. Effects of simultaneous calcination and reduction on performance of promoted Ni/SiO2 catalyst in steam reforming of propane. Int. J. Hydrogen Energy 2019, 44, 9307–9315. [Google Scholar] [CrossRef]
  16. Harshini, D.; Yoon, C.W.; Han, J.; Yoon, S.P.; Nam, S.W.; Lim, T.-H. Catalytic Steam Reforming of Propane over Ni/LaAlO3 Catalysts: Influence of Preparation Methods and OSC on Activity and Stability. Catal. Lett. 2012, 142, 205–212. [Google Scholar] [CrossRef]
  17. Azizzadeh Fard, A.; Arvaneh, R.; Alavi, S.M.; Bazyari, A.; Valaei, A. Propane steam reforming over promoted Ni–Ce/MgAl2O4 catalysts: Effects of Ce promoter on the catalyst performance using developed CCD model. Int. J. Hydrogen Energy 2019, 44, 21607–21622. [Google Scholar] [CrossRef]
  18. Kim, K.M.; Kwak, B.S.; Park, N.-K.; Lee, T.J.; Lee, S.T.; Kang, M. Effective hydrogen production from propane steam reforming over bimetallic co-doped NiFe/Al2O3 catalyst. J. Ind. Eng. Chem. 2017, 46, 324–336. [Google Scholar] [CrossRef]
  19. Laosiripojana, N.; Sangtongkitcharoen, W.; Assabumrungrat, S. Catalytic steam reforming of ethane and propane over CeO2-doped Ni/Al2O3 at SOFC temperature: Improvement of resistance toward carbon formation by the redox property of doping CeO2. Fuel 2006, 85, 323–332. [Google Scholar] [CrossRef]
  20. Natesakhawat, S.; Oktar, O.; Ozkan, U.S. Effect of lanthanide promotion on catalytic performance of sol–gel Ni/Al2O3 catalysts in steam reforming of propane. J. Mol. Catal. A Chem. 2005, 241, 133–146. [Google Scholar] [CrossRef]
  21. Matsumura, Y.; Nakamori, T. Steam reforming of methane over nickel catalysts at low reaction temperature. Appl. Catal. A Gen. 2004, 258, 107–114. [Google Scholar] [CrossRef]
  22. Guo, J.; Lou, H.; Zhao, H.; Chai, D.; Zheng, X. Dry reforming of methane over nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A Gen. 2004, 273, 75–82. [Google Scholar] [CrossRef]
  23. Lee, D.S.; Min, D.J. A Kinetics of Hydrogen Reduction of Nickel Oxide at Moderate Temperature. Met. Mater. Int. 2019, 25, 982–990. [Google Scholar] [CrossRef] [Green Version]
  24. Richardson, J.T.; Scates, R.; Twigg, M.V. X-ray diffraction study of nickel oxide reduction by hydrogen. Appl. Catal. A Gen. 2003, 246, 137–150. [Google Scholar] [CrossRef]
  25. Richardson, J.T.; Scates, R.M.; Twigg, M.V. X-ray diffraction study of the hydrogen reduction of NiO/α-Al2O3 steam reforming catalysts. Appl. Catal. A Gen. 2004, 267, 35–46. [Google Scholar] [CrossRef]
  26. Almasan, V.; Gaeumann, T.; Lazar, M.; Marginean, P.; Aldea, N. Hydrogen spillover effect over the oxide surfaces in supported nickel catalysts. In Studies in Surface Science and Catalysis; Froment, G.F., Waugh, K.C., Eds.; Elsevier: Amsterdam, The Netherlands, 1997; Volume 109, pp. 547–552. [Google Scholar]
  27. Kramer, R.; Andre, M. Adsorption of atomic hydrogen on alumina by hydrogen spillover. J. Catal. 1979, 58, 287–295. [Google Scholar] [CrossRef]
  28. Hatzisymeon, M.; Petala, A.; Panagiotopoulou, P. Carbon Dioxide Hydrogenation over Supported Ni and Ru Catalysts. Catal. Lett. 2021, 151, 888–900. [Google Scholar] [CrossRef]
  29. Kolb, G.; Zapf, R.; Hessel, V.; Löwe, H. Propane steam reforming in micro-channels—results from catalyst screening and optimisation. Appl. Catal. A Gen. 2004, 277, 155–166. [Google Scholar] [CrossRef]
  30. Modafferi, V.; Panzera, G.; Baglio, V.; Frusteri, F.; Antonucci, P.L. Propane reforming on Ni–Ru/GDC catalyst: H2 production for IT-SOFCs under SR and ATR conditions. Appl. Catal. A Gen. 2008, 334, 1–9. [Google Scholar] [CrossRef]
  31. Malaibari, Z.O.; Croiset, E.; Amin, A.; Epling, W. Effect of interactions between Ni and Mo on catalytic properties of a bimetallic Ni-Mo/Al2O3 propane reforming catalyst. Appl. Catal. A Gen. 2015, 490, 80–92. [Google Scholar] [CrossRef]
  32. Helali, Z.; Jedidi, A.; Syzgantseva, O.A.; Calatayud, M.; Minot, C. Scaling reducibility of metal oxides. Theor. Chem. Acc. 2017, 136, 100–115. [Google Scholar] [CrossRef]
  33. Alphonse, P.; Ansart, F. Catalytic coatings on steel for low-temperature propane prereforming to solid oxide fuel cell (SOFC) application. J. Colloid Interface Sci. 2009, 336, 658–666. [Google Scholar] [CrossRef] [Green Version]
  34. Panagiotopoulou, P.; Kondarides, D.I.; Verykios, X.E. Mechanistic aspects of the selective methanation of CO over Ru/TiO2 catalyst. Catal. Today 2012, 181, 138–147. [Google Scholar] [CrossRef]
  35. Panagiotopoulou, P. Methanation of CO2 over alkali-promoted Ru/TiO2 catalysts: II. Effect of alkali additives on the reaction pathway. Appl. Catal. B Environ. 2018, 236, 162–170. [Google Scholar] [CrossRef]
  36. Wang, B.; Wu, X.; Ran, R.; Si, Z.; Weng, D. IR characterization of propane oxidation on Pt/CeO2–ZrO2: The reaction mechanism and the role of Pt. J. Mol. Catal. A Chem. 2012, 356, 100–105. [Google Scholar] [CrossRef]
  37. Panagiotopoulou, P.; Christodoulakis, A.; Kondarides, D.I.; Boghosian, S. Particle size effects on the reducibility of titanium dioxide and its relation to the water–gas shift activity of Pt/TiO2 catalysts. J. Catal. 2006, 240, 114–125. [Google Scholar] [CrossRef]
  38. Kokka, A.; Ramantani, T.; Petala, A.; Panagiotopoulou, P. Effect of the nature of the support, operating and pretreatment conditions on the catalytic performance of supported Ni catalysts for the selective methanation of CO. Catal. Today 2020, 355, 832–843. [Google Scholar] [CrossRef]
  39. Karelovic, A.; Ruiz, P. Mechanistic study of low temperature CO2 methanation over Rh/TiO2 catalysts. J. Catal. 2013, 301, 141–153. [Google Scholar] [CrossRef]
  40. Liao, L.F.; Lien, C.F.; Shieh, D.L.; Chen, M.T.; Lin, J.L. FTIR Study of Adsorption and Photoassisted Oxygen Isotopic Exchange of Carbon Monoxide, Carbon Dioxide, Carbonate, and Formate on TiO2. J. Phys. Chem. B 2002, 106, 11240–11245. [Google Scholar] [CrossRef]
  41. Marwood, M.; Doepper, R.; Renken, A. In-situ surface and gas phase analysis for kinetic studies under transient conditions The catalytic hydrogenation of CO2. Appl. Catal. A Gen. 1997, 151, 223–246. [Google Scholar] [CrossRef] [Green Version]
  42. Mino, L.; Spoto, G.; Ferrari, A.M. CO2 Capture by TiO2 Anatase Surfaces: A Combined DFT and FTIR Study. J. Phys. Chem. C 2014, 118, 25016–25026. [Google Scholar] [CrossRef]
  43. Aldana, P.A.U.; Ocampo, F.; Kobl, K.; Louis, B.; Thibault-Starzyk, F.; Daturi, M.; Bazin, P.; Thomas, S.; Roger, A.C. Catalytic CO2 valorization into CH4 on Ni-based ceria-zirconia. Reaction mechanism by operando IR spectroscopy. Catal. Today 2013, 215, 201–207. [Google Scholar] [CrossRef]
  44. Westermann, A.; Azambre, B.; Bacariza, M.C.; Graça, I.; Ribeiro, M.F.; Lopes, J.M.; Henriques, C. Insight into CO2 methanation mechanism over NiUSY zeolites: An operando IR study. Appl. Catal. B Environ. 2015, 174–175, 120–125. [Google Scholar] [CrossRef]
  45. Mohamed, Z.; Dasireddy, V.D.B.C.; Singh, S.; Friedrich, H.B. The preferential oxidation of CO in hydrogen rich streams over platinum doped nickel oxide catalysts. Appl. Catal. B Environ. 2016, 180, 687–697. [Google Scholar] [CrossRef]
  46. Konishcheva, M.V.; Potemkin, D.I.; Badmaev, S.D.; Snytnikov, P.V.; Paukshtis, E.A.; Sobyanin, V.A.; Parmon, V.N. On the Mechanism of CO and CO2 Methanation Over Ni/CeO2 Catalysts. Top. Catal. 2016, 59, 1424–1430. [Google Scholar] [CrossRef]
  47. Köck, E.-M.; Kogler, M.; Bielz, T.; Klötzer, B.; Penner, S. In Situ FT-IR Spectroscopic Study of CO2 and CO Adsorption on Y2O3, ZrO2, and Yttria-Stabilized ZrO2. J. Phys. Chem. C 2013, 117, 17666–17673. [Google Scholar] [CrossRef]
  48. Bachiller-Baeza, B.; Rodriguez-Ramos, I.; Guerrero-Ruiz, A. Interaction of Carbon Dioxide with the Surface of Zirconia Polymorphs. Langmuir 1998, 14, 3556–3564. [Google Scholar] [CrossRef]
  49. Zhang, Z.; Zhang, L.; Hülsey, M.J.; Yan, N. Zirconia phase effect in Pd/ZrO2 catalyzed CO2 hydrogenation into formate. Mol. Catal. 2019, 475, 110461. [Google Scholar] [CrossRef]
  50. Föttinger, K.; Emhofer, W.; Lennon, D.; Rupprechter, G. Adsorption and Reaction of CO on (Pd–)Al2O3 and (Pd–)ZrO2: Vibrational Spectroscopy of Carbonate Formation. Topics Catal. 2017, 60, 1722–1734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Pokrovski, K.; Jung, K.T.; Bell, A.T. Investigation of CO and CO2 Adsorption on Tetragonal and Monoclinic Zirconia. Langmuir 2001, 17, 4297–4303. [Google Scholar] [CrossRef]
  52. Anderson, J.A.; Daza, L.; Fierro, J.L.G.; Rodrigo, M.T. Influence of preparation method on the characteristics of nickel/sepiolite catalysts. J. Chem. Soc. Faraday Trans. 1993, 89, 3651–3657. [Google Scholar] [CrossRef]
  53. Wu, R.-f.; Zhang, Y.; Wang, Y.-z.; Gao, C.-g.; Zhao, Y.-x. Effect of ZrO2 promoter on the catalytic activity for CO methanation and adsorption performance of the Ni/SiO2 catalyst. J. Fuel Chem. Technol. 2009, 37, 578–582. [Google Scholar] [CrossRef]
  54. Zhou, G.; Liu, H.; Cui, K.; Jia, A.; Hu, G.; Jiao, Z.; Liu, Y.; Zhang, X. Role of surface Ni and Ce species of Ni/CeO2 catalyst in CO2 methanation. Appl. Surf. Sci. 2016, 383, 248–252. [Google Scholar] [CrossRef]
  55. Faria, E.C.; Rabelo-Neto, R.C.; Colman, R.C.; Ferreira, R.A.R.; Hori, C.E.; Noronha, F.B. Steam Reforming of LPG over Ni/Al2O3 and Ni/CexZr1—xO2/Al2O3 Catalysts. Catal. Lett. 2016, 146, 2229–2241. [Google Scholar] [CrossRef]
Figure 1. XRD patterns obtained from 5 wt.% Ni catalysts supported on the indicated commercial metal oxide carriers. The reflection planes of anatase (°) and rutile (*) TiO2 phases are indicated in the diffractogram of Ni/TiO2 catalyst.
Figure 1. XRD patterns obtained from 5 wt.% Ni catalysts supported on the indicated commercial metal oxide carriers. The reflection planes of anatase (°) and rutile (*) TiO2 phases are indicated in the diffractogram of Ni/TiO2 catalyst.
Nanomaterials 11 01948 g001
Figure 2. TEM images and Selected Area Electron Diffraction (SAED) patterns obtained for (A) 5%Ni/YSZ, (B) 5%Ni/CeO2 and (C) 5%Ni/TiO2 catalysts. Ni particles are indicated with red arrows.
Figure 2. TEM images and Selected Area Electron Diffraction (SAED) patterns obtained for (A) 5%Ni/YSZ, (B) 5%Ni/CeO2 and (C) 5%Ni/TiO2 catalysts. Ni particles are indicated with red arrows.
Nanomaterials 11 01948 g002
Figure 3. (A) Conversions of C3H8 as a function of reaction temperature and (B) Arrhenius plots of turnover frequencies of C3H8 conversion obtained over Ni catalysts (5.0 wt.%) supported on the indicated commercial oxide carriers. Experimental conditions: Mass of catalyst: 150 mg; particle diameter: 0.15 < dp < 0.25 mm; Feed composition: 4.5% C3H8, 0.15% Ar, 44% H2O (balance He); Total flow rate: 250 cm3 min−1.
Figure 3. (A) Conversions of C3H8 as a function of reaction temperature and (B) Arrhenius plots of turnover frequencies of C3H8 conversion obtained over Ni catalysts (5.0 wt.%) supported on the indicated commercial oxide carriers. Experimental conditions: Mass of catalyst: 150 mg; particle diameter: 0.15 < dp < 0.25 mm; Feed composition: 4.5% C3H8, 0.15% Ar, 44% H2O (balance He); Total flow rate: 250 cm3 min−1.
Nanomaterials 11 01948 g003
Figure 4. Selectivities toward reaction products as a function of reaction temperature obtained over Ni catalysts (5.0 wt.%) supported on (A) ZrO2, (B) YSZ, (C) Al2O3, (D) CeO2 and (E) TiO2. Experimental conditions: same as in Figure 3.
Figure 4. Selectivities toward reaction products as a function of reaction temperature obtained over Ni catalysts (5.0 wt.%) supported on (A) ZrO2, (B) YSZ, (C) Al2O3, (D) CeO2 and (E) TiO2. Experimental conditions: same as in Figure 3.
Nanomaterials 11 01948 g004
Figure 5. Turnover frequencies of C3H8 conversion as a function of selectivity toward CH4 obtained at 450 °Cover supported Ni catalysts.
Figure 5. Turnover frequencies of C3H8 conversion as a function of selectivity toward CH4 obtained at 450 °Cover supported Ni catalysts.
Nanomaterials 11 01948 g005
Figure 6. Long-term stability test of the 5%Ni/ZrO2 catalyst at 650 °C: Alterations of the conversion of C3H8 and selectivities toward reaction products with time-on-stream. Experimental conditions: Same as in Figure 3. Dashed vertical black lines indicate shutting down of the system overnight.
Figure 6. Long-term stability test of the 5%Ni/ZrO2 catalyst at 650 °C: Alterations of the conversion of C3H8 and selectivities toward reaction products with time-on-stream. Experimental conditions: Same as in Figure 3. Dashed vertical black lines indicate shutting down of the system overnight.
Nanomaterials 11 01948 g006
Figure 7. DRIFT spectra obtained over the (A) Ni/TiO2 and (B) Ni/ZrO2 catalysts following interaction with 0.5% C3H8 + 5% H2O (in He) at 100 °C for 15 min and subsequent stepwise heating at 500 °C.
Figure 7. DRIFT spectra obtained over the (A) Ni/TiO2 and (B) Ni/ZrO2 catalysts following interaction with 0.5% C3H8 + 5% H2O (in He) at 100 °C for 15 min and subsequent stepwise heating at 500 °C.
Nanomaterials 11 01948 g007
Figure 8. DRIFT spectra obtained in the region 3200–2400 cm−1 over the (A) Ni/TiO2 and (B) Ni/ZrO2 catalysts following interaction with 0.5% C3H8 + 5% H2O (in He) at 100 °C and 450 °C for 15 min.
Figure 8. DRIFT spectra obtained in the region 3200–2400 cm−1 over the (A) Ni/TiO2 and (B) Ni/ZrO2 catalysts following interaction with 0.5% C3H8 + 5% H2O (in He) at 100 °C and 450 °C for 15 min.
Nanomaterials 11 01948 g008
Table 1. Physicochemical properties of supported Ni (5 wt.%) catalysts and their apparent activation energies for propane steam reforming reaction.
Table 1. Physicochemical properties of supported Ni (5 wt.%) catalysts and their apparent activation energies for propane steam reforming reaction.
CatalystSSA (a) (m2/g)dMxOy (b) (nm)DNi (c) (%)dNi (c) (nm)Activation Energy (kJ/mol)
5%Ni/ZrO23915.05.717.8154
5%Ni/YSZ1120.94.721.4140
5%Ni/Al2O3666.04.025.5121
5%Ni/CeO23910.511.98.5102
5%Ni/TiO24121.82.836.1127
(a) Specific surface area, estimated with the BET method. (b) Primary crystallite size of MxOy, estimated from XRD line broadening. (c) Dispersion and mean particle size of Ni, estimated from selective chemisorption measurements.
Table 2. Selected area electron diffraction (SAED) analysis of TEM images obtained for 5%Ni/YSZ, 5%Ni/CeO2 and 5%Ni/TiO2 catalysts.
Table 2. Selected area electron diffraction (SAED) analysis of TEM images obtained for 5%Ni/YSZ, 5%Ni/CeO2 and 5%Ni/TiO2 catalysts.
CatalystSpotd-Spacing (Å)FormulaCrystallographic Plane (h k l)JCPDS No
5%Ni/YSZ13.0Y0.15Zr0.85O1.93(111)30-1468
22.6Y0.15Zr0.85O1.93(200)30-1468
32.1Ni(111)1-1258
5%Ni/CeO213.1CeO2(111)1-800
22.7CeO2(200)1-800
32.1Ni(111)1-1258
5%Ni/TiO213.5TiO2(101)1-562
22.4TiO2(103)1-562
32.1Ni(111)1-1258
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kokka, A.; Petala, A.; Panagiotopoulou, P. Support Effects on the Activity of Ni Catalysts for the Propane Steam Reforming Reaction. Nanomaterials 2021, 11, 1948. https://doi.org/10.3390/nano11081948

AMA Style

Kokka A, Petala A, Panagiotopoulou P. Support Effects on the Activity of Ni Catalysts for the Propane Steam Reforming Reaction. Nanomaterials. 2021; 11(8):1948. https://doi.org/10.3390/nano11081948

Chicago/Turabian Style

Kokka, Aliki, Athanasia Petala, and Paraskevi Panagiotopoulou. 2021. "Support Effects on the Activity of Ni Catalysts for the Propane Steam Reforming Reaction" Nanomaterials 11, no. 8: 1948. https://doi.org/10.3390/nano11081948

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