Next Article in Journal
Effect of Silica Embedding on the Structure, Morphology and Magnetic Behavior of (Zn0.6Mn0.4Fe2O4)δ/(SiO2)(100−δ) Nanoparticles
Next Article in Special Issue
PbS Quantum Dots Saturable Absorber for Dual-Wavelength Solitons Generation
Previous Article in Journal
Effect of Wettability and Uniform Distribution of Reinforcement Particle on Mechanical Property (Tensile) in Aluminum Metal Matrix Composite—A Review
Previous Article in Special Issue
Evaluation of Local Mechanical and Chemical Properties via AFM as a Tool for Understanding the Formation Mechanism of Pulsed UV Laser-Nanoinduced Patterns on Azo-Naphthalene-Based Polyimide Films
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Promoting the Selectivity of Pt/m-ZrO2 Ethanol Steam Reforming Catalysts with K and Rb Dopants

1
Center for Applied Energy Research, University of Kentucky, 2540 Research Park Drive, Lexington, KY 40511, USA
2
Department of Biomedical Engineering and Chemical Engineering, The University of Texas at San Antonio, 1 UTSA Circle, San Antonio, TX 78249, USA
3
Argonne National Laboratory, Lemont, IL 60439, USA
4
Department of Mechanical Engineering, The University of Texas at San Antonio, 1 UTSA Circle, San Antonio, TX 78249, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2233; https://doi.org/10.3390/nano11092233
Submission received: 30 July 2021 / Revised: 25 August 2021 / Accepted: 25 August 2021 / Published: 29 August 2021

Abstract

:
The ethanol steam reforming reaction (ESR) was investigated on unpromoted and potassium- and rubidium-promoted monoclinic zirconia-supported platinum (Pt/m-ZrO2) catalysts. Evidence from in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) characterization indicates that ethanol dissociates to ethoxy species, which undergo oxidative dehydrogenation to acetate followed by acetate decomposition. The acetate decomposition pathway depends on catalyst composition. The decarboxylation pathway tends to produce higher overall hydrogen selectivity and is the most favored route at high alkali loading (2.55 wt.% K and higher or 4.25 wt.% Rb and higher). On the other hand, decarbonylation is a significant route for the undoped catalyst or when a low alkali loading (e.g., 0.85% K or 0.93% Rb) is used, thus lowering the overall H2 selectivity of the process. Results of in situ DRIFTS and the temperature-programmed reaction of ESR show that alkali doping promotes forward acetate decomposition while exposed metallic sites tend to facilitate decarbonylation. In previous work, 1.8 wt.% Na was found to hinder decarbonylation completely. Due to the fact that 1.8 wt.% Na is atomically equivalent to 3.1 wt.% K and 6.7 wt.% Rb, the results show that less K (2.55% K) or Rb (4.25% Rb) is needed to suppress decarbonylation; that is, more basic cations are more efficient promoters for improving the overall hydrogen selectivity of the ESR process.

1. Introduction

In recent decades, the catalytic steam reforming of hydrocarbons such as natural gas has been the most economically competitive method to produce hydrogen in the chemical industry. However, this method is not sustainable as the feedstock is a fossil source, and significant amounts of CO2 are produced in the process. Thus, the development of new sustainable reforming technologies from renewable feedstocks (e.g., biomass-derived oxygenates) is necessary for reducing net greenhouse gas emissions [1]. In this scenario, researchers are focusing on several renewable feedstocks such as ethanol, polyols, and dimethyl ether [2,3,4,5]. Among these renewable feedstocks, bio-ethanol is very attractive because of its favorable hydrogen content, wide abundance, low toxicity, and ability to be easily stored for transportation and portable power [3]. Moreover, while ethanol production (e.g., from sugar cane or corn) currently competes with food production, cellulosic ethanol is currently under development. The overall ethanol steam reforming (ESR) reaction, which occurs at 350–650 °C, can be summarized by Equation (1):
C2H5OH + 3H2O → 2CO2 + 6H2
Acetaldehyde, ethylene, and methane are formed during ESR. However, the concentration of these species must be minimized to achieve higher efficiency of hydrogen production and avoid carbon formation and consequently catalyst deactivation [6]. These intermediates and byproducts can be converted via different pathways, depending on catalyst structure as well as the conditions used in the reactor [3].
Ethanol steam reforming catalysts are often transition metals such as copper, cobalt, and nickel, or noble metals such as platinum, palladium, rhodium, gold, or ruthenium; combinations of metals have also been used [7,8]. Among the transition metals, nickel and cobalt are the most common. However, catalyst deactivation by coke deposition is a major issue for these catalysts [6,8,9,10,11,12,13,14]. Noble metals such as Pt and Rh have better resistance to coke deposition as well as high activity, but they are much higher in cost [8,15,16,17,18]. These metals are typically supported on basic, acidic, or inert supports [7]. Oxides that form surface defects through partial reduction (e.g., ceria, zirconia, ceria–zirconia mixtures, and metal-promoted ceria catalysts) have also been investigated because of their high oxygen mobility, their ability to dissociate water or ROH (i.e., where R is an alkyl group - methanol, ethanol, etc.), and their capability to shuttle O-bound intermediates on the catalyst surface [19,20]. Ciambelli et al. [21] observed that Pt/CeO2 exhibited higher activity as compared to Pt/Al2O3 for ESR in the temperature range of 300–450 °C. He et al. [22] obtained similar findings. The high activity of Pt/CeO2 was due in part to the ability of Pt to facilitate the formation of defect sites on the partially reducible oxide support. ROH molecules can then dissociate on these defect sites [23], which is comparable to the adsorption of H2O at defect sites that result in the formation of bridging OH groups [24,25,26].
Recently, the alkali promotion of noble or transition metal catalysts has been investigated. Alkali doping can impact many catalyst properties, including activity, stability, resistance to coke formation, selectivity, and surface acidity/basicity [6,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. The effect of alkali on nickel-based catalysts is contradictory. Frusteri et al. [33,35] found that Li and Na adversely affect the nickel dispersion, whereas they improve the extent of reduction of nickel. In contrast, no effect on either Ni dispersion or catalyst morphology was detected for K-promoted catalysts. Moreover, the authors examined the influence of alkali loading on catalyst performance. Lithium and potassium enhanced catalyst stability mainly by depressing nickel sintering, whereas carbon laydown did not seem to be influenced by adding K. However, a different effect of potassium was observed by Slowik et al. [6]. In that case, potassium promotion of Ni/CeO2 was not found to protect the catalyst against the formation of carbon deposits and did not improve stability during ESR.
Improvements in stability/activity were reported for cobalt-based catalysts, and this is mainly related to the inhibition of carbon deposition [27,28,29,30,31,38]. Recently, Grzybek et al. [27] studied the alkali surface state (location, dynamics) by Species Resolved Thermal Alkali Desorption (SR-TAD). Movement of potassium from cobalt to alumina was observed during both activation and ESR. This phenomenon stabilizes small cobalt crystallites by hindering their detachment from the catalyst surface, which would otherwise result in encapsulation by the growing carbonaceous deposit. Furthermore, Grzybek et al. [28] found that potassium loadings from 0.1 to 4 wt.% improve the activity of their catalyst by enhancing C–C bond scission, but 0.3 wt.% is the optimal loading to maximize the selectivity to H2 and CO2, the most desirable products. The beneficial effect of K can also be related to improvement in reduction of Co2+ to Co0, the stabilization of acetate species, and the suppression of methane formation [31].
To our knowledge, not many studies are available on the effect of alkali for noble metal catalysts, especially at low temperatures [34,40,41,42]. Low potassium loading (0.2%) was found to decrease the initial conversion but improves the stability for Rh/CeO2-ZrO2. At higher loading (5%), the catalyst activity is negligible [42]. Dömök et al. explored doping potassium to Pt/Al2O3 catalyst, and they found that increasing potassium content progressively decreases the ethanol conversion and changes the product distribution toward a higher selectivity to CH4 and CO2 as compared to 1% Pt/Al2O3 [34]. Furthermore, potassium destabilizes adsorbed acetate, promoting its decomposition to CO2 and CH4 at a lower temperature; on the other hand, acetate species were more stable on the undoped catalyst, decomposing at ~420 °C [34].
In our prior study of Na-doping to Pt/ZrO2 [40,41], a similar trend was obtained, as acetate decomposed at a lower temperature (100–150 °C) when 1.8–2.5 wt.% Na was added to the formulation as compared to the undoped catalyst. Our focus is on the low-temperature conversion of ethanol with steam to H2, CO2, and CH4, with the latter being reformed in a conventional methane reforming (e.g., autothermal reforming). With this decarboxylation route Equation (1), which involves the steps below Equations (2) and (3), higher hydrogen selectivity is expected as compared to the decarbonylation route Equation (4):
Decarboxylation pathway:
C2H5OH + H2O → 2H2 + CH4 + CO2
CH4 + 2H2O → 4H2 + CO2
Decarbonylation pathway
C2H5OH + H2O → 4H2 + 2CO
In that work, DRIFTS experiments revealed that the acetate decomposition pathway depends on the Na loading. Forward direct acetate decomposition to CH4 and carbonate Equation (2) is the most favorable pathway at high sodium loading (1.8 or 2.5 wt.%), whereas the unselective decarbonylation route occurs for the unpromoted catalyst or at low sodium loading (0.5 wt.%) and is promoted by metallic sites.
The question remains as to whether more basic alkali metals might further improve the decarboxylation selectivity over that of decarbonylation during ESR. To that end, the effect of potassium and rubidium loading on the relative rates of decarboxylation/decarbonylation was investigated. Pt/m-ZrO2 catalyst was promoted by the following potassium loadings: 0% (reference), 0.85%, 1.70%, 2.55%, 3.40%, 4.25%, and 8.50 wt.%; whereas rubidium was added with the following loadings: 0.93%, 1.86%, 2.79%, 3.72%, 4.65%, 5.59%, and 9.29%. Atomically equivalent loadings allowed for some comparisons between the K- and Rb-promoted catalysts, as well as with catalysts prepared in our earlier study using Na as the dopant [40,41]. These systems were characterized by N2 physisorption, transmission electron microscopy (TEM), X-ray absorption near edge spectroscopy (XANES), extended X-ray absorption fine structure spectroscopy (EXAFS), hydrogen temperature programmed reduction (H2-TPR), temperature-programmed ESR, and DRIFTS. Catalyst activity and selectivity were measured at steady-state using a fixed bed tubular reactor.

2. Materials and Methods

2.1. Catalyst Preparation

Various potassium and rubidium loadings on 2% Pt/ZrO2 were prepared via incipient wetness impregnation (IWI). Firstly, monoclinic ZrO2 (Alfa Aesar, Haverhill, MA, USA) was impregnated by 2% Pt with an aqueous solution of tetraamine platinum (II) nitrate (Alfa Aesar, Haverhill, MA, USA) followed by drying and calcination at 350 °C (four hours, muffle furnace). Then, the appropriate amount of KNO3 (Alfa Aesar, Haverhill, MA, USA) or RbNO3 (Alfa Aesar, Haverhill, MA, USA) was added via IWI. Promoted catalysts were dried and recalcined using the same conditions.

2.2. Catalyst Characterization

Surface area and pore size were determined using a Micromeritics 3-Flex instrument (Micromeritics, Norcross, GA, USA). The BJH method was used to calculate the average pore diameter and specific volume. Samples were pre-treated at 160 °C at 50 mTorr for no less than 12 h.
H2 temperature-programmed reduction (TPR) plots of the catalysts were obtained using an Altamira AMI-300R (Altamira, Pittsburgh, PA, USA) instrument employing a thermal conductivity detector (TCD). During experiments, approximately 200 mg of catalyst was loaded into the U-tube reactor, and then a mixture of 10% H2 in Ar (30 cm3/min) (Airgas, San Antonio, TX, USA) was flowed while the temperature was ramped from 50 to 1000 °C at a heating rate of 10 °C/min.
Transmission electron microscopy (TEM) and scanning transmission microscopy (STEM) were conducted with an FEI Talos F200X scope (Thermo Fisher Scientific, Waltham, MA, USA) equipped with bright field (BF), dark-field (DF) 2, DF 4, and high-angle annular dark-field (HAADF) detectors. Imaging was performed using a field emission gun with an accelerating voltage of 200 kV and a high-speed Ceta 16M camera. The elemental distributions were determined via FEI super energy-dispersive X-ray spectroscopy (EDX) (Thermo Fisher Scientific, Waltham, MA, USA). Velox software was utilized to process the data. Before analysis, samples were first treated in H2 at 350 °C for 1 hour and then cooled to room temperature, followed by passivation with 1% O2 (balance N2). Then, the reduced/passivated samples were sonicated for 30 min in ethanol. A sample of this suspension was dropped onto a lacey carbon-covered Cu grid (300 mesh) and dried in air for 12 h.
Temperature-programmed reaction/desorption analyses were carried out with the Altamira AMI-300R unit. For each experiment, the catalyst was reduced at 300 °C using 10 cm3/min H2 (Airgas, San Antonio, TX, USA) and 20 cm3/min argon (Airgas, San Antonio, TX, USA) for 1 h. After cooling the catalyst to 50 °C in flowing Ar, ethanol was pumped at a rate of 100 mL/min across the catalyst for 10 min, and then Ar was flowed at 30 cm3/min for 20 min. Following this purge, water was bubbled in He at a rate of 30 cm3/min for 10 min, and then another purge in Ar was performed. Finally, Ar was flowed at 30 cm3/min while the temperature was increased to 700 °C. This allowed analysis of ESR under transient conditions. CH4 evolution was measured to examine the effect of K or Rb on C–C cleaving.
A Nicolet (Thermo Fisher, Waltham, MA, USA) iS-10 Fourier Transform infrared spectrometer, coupled with a Harrick Scientific (Pleasantville, New York, NY, USA) Praying Mantis accessory, was used for the temperature desorption/reaction experiments. The catalyst was reduced at 300 °C with a 1:1 mixture of H2:He at a flow rate of 200 cm3/min for 1 h and then cooled to 50 °C in hydrogen. Next, helium was used to bubble ethanol at a flow rate of 75 cm3/min for 15 min. Subsequently, He was bubbled through water (31 °C water bath), giving an H2O concentration of 4.4% with a flow rate of 75 cm3/min. Temperature was stepped in 50 °C increments from 50 °C to 500 °C. At each step, 512 scans were taken at a resolution of 4.
Temperature programmed reduction with X-ray absorption near edge spectroscopy (TPR-XANES) was performed at the Materials Research Collaborative Access Team (MR-CAT) beamline located at the Advanced Photon Source, Argonne National Laboratory. Sample amount was optimized for the platinum L3 edge. The quartz reactor was held in a clamshell furnace located on a positioning table, and the beam passed through six samples in a sequential manner with 20 μm accuracy for repeat scans. He was flowed through the catalysts for more than 5 min at a flow rate of 30 mL/min. Pure hydrogen was then passed through the sample at a flow rate of 30 mL/min, and a heating rate of 0.83 °C/min was started for the furnace to achieve 300 °C. After soaking at this temperature for 1 h, samples were cooled to room temperature and scanned for both the XANES and EXAFS regions at both the platinum L3 (11.564 keV) and L2 edges (13.273 keV), so that the L3-L2 edge difference procedure could be applied [43]. Spectra were recorded in transmission mode, and the respective metallic foil was measured in a concurrent manner for the purpose of energy calibration. The Pt L3 edge data were in the range of 11.400–12.700 keV, and Pt L2 edge data were in the range of 13.100–13.850 keV. Standard data reduction was conducted with WinXAS (Version 2.0, Thorsten Ressler, Berlin, Germany) [44], while fittings were carried out for EXAFS with Atoms (Copyright 2001, Department of Physics, University of Washington, Seattle, WA, USA) [45], FEFF8 (Version 8.20, Department of Physics, University of Washington, Seattle, WA, USA) [46], and FEFFIT (Copyright 2001, Department of Physics, University of Washington, Seattle, WA, USA) [46] software over Δk = 3–10 Å−1 and ΔR = 1.85–3.25 Å. A typical analysis included post-edge background subtraction (Victoreen function), pre-edge and post-edge background subtraction (degree 1 polynomials), normalization based on the edge jump, conversion to k-space with background subtraction using a cubic spline, and applying a Fourier transform of the χ(k) function to R-space.

2.3. Catalytic Activity

The activity of the catalysts was tested in a fixed bed reactor (stainless steel tubular reactor, I.D. 0.444 in.); more information on the experimental set-up is reported in our previous study [47]. Briefly, 80 mg of catalyst (63–106 µm) was diluted with 300 mg of SiO2 beads and activated using 100 cm3/min H2 at 350 °C for 1 h. Next, the temperature was cooled to 300 °C, and the gas was changed to a mixture containing 26.1% H2O, 2.9% C2H5OH (balance N2) at P = 1 atm, gas hourly space velocity (GHSV) = 190,560 Ncm3/min/gcat. The unpromoted catalyst was also tested at different GHSV in order to compare the selectivities among the catalysts at similar conversion. The products were then passed through a cold trap (held at 5 °C) to collect condensable compounds. The condensable products were analyzed by SRI (SRI Instruments, Torrance, CA, USA) 8610 GC equipped with HayeSep Q-column, whereas the gas products were analyzed by Inficon micro-GC Fusion equipped by molecular sieve, alumina, plot-u, and OV-1. Ethanol conversion Equation (5), carbon selectivity Equation (6), and H2 yield Equation (7) were calculated using the following formulas:
χ C 2 H 5 O H = 1 F C 2 H 5 O H o u t F C 2 H 5 O H i n
S i = n i · F i o u t ,   p r o d i n i · F i o u t , p r o d
H 2   y i e l d = F H 2 o u t ,   p r o d 6 · F C 2 H 5 O H i n
where F C 2 H 5 O H i n is the molar feed rate of ethanol, F C 2 H 5 O H o u t is the effluent molar flow rate, n i is carbon number, F H 2 o u t ,   p r o d   is the effluent molar flow rate of H2, and F i o u t ,   p r o d is effluent molar flow rate of the C-containing species (i = CO, CO2, C2H6, C2H4, C3H6, C2H4O).

3. Results and Discussion

Surface area and porosity results for un-promoted and K and Rb-promoted catalysts are provided in Table 1. Both K and Rb series showed that low alkali loadings caused a slight decrease in the surface area, whereas high loadings dramatically decreased it (i.e., below what is expected from the decrease due to the added mass of alkali). For example, the surface area dropped from 89.7 m2/gcat (unpromoted) to 34.7 and 58.2 m2/gcat for 8.5% K and 9.29% Rb, respectively. This decrease in the surface area suggests that pore blocking is more significant at high alkali loading. Increasing alkali doping progressively diminished the pore volume, but little impact was observed for the average pore diameter of the Rb-promoted catalyst. In contrast, above 2.55% K, the average diameter increased systematically, suggesting preferential blocking of narrower pores by K at higher loadings.
Estimated Pt diameter and dispersion obtained by EXAFS fittings are also reported in Table 1. Interestingly, the diameter of the Pt cluster increases with the alkali loading, and the effect is more pronounced for the K-promoted catalyst. This results in a decreasing trend in Pt dispersion, which drops from 88% and 94% for the unpromoted catalysts to 35% and 56% for 4.25% K-Pt/ZrO2 and 9.29% Rb-Pt/ZrO2, respectively.
STEM-EDX images for Rb-promoted catalysts are shown in Figure 1. Both platinum and rubidium were well dispersed for 0.93% Rb-2% Pt/ZrO2, and all clusters observed were below 3 nm. By increasing the rubidium loading to 9.29%, STEM-EDX images (Figure 1, bottom) show rubidium was well dispersed, whereas platinum particles tended to form agglomerates of several Pt domains. Some agglomerates were on the order of 10 nm, while the domains were typically below 3 nm. The spatial distribution of Pt and Rb also suggests, especially in the case of 9.29% Rb, that there is a strong possibility of contact between the two elements.
Catalyst reducibility was investigated by TPR (Figure 2). The unpromoted Pt/zirconia catalyst had a small hydrogen uptake in the range of 150–200 °C, indicative of reduction of platinum oxide to metal, which occurs around 200 °C, and Pt-catalyzed defect formation [46]. Pt accelerates the decomposition of surface carbonates and facilitates the formation of oxygen vacancies and bridging OH groups [25,52,53]. The hydrogen uptake during TPR further increased by adding potassium or rubidium.
Temperature programmed reduction with mass spectrometry (TPR-MS) spectra for K and Rb-promoted catalysts, reported in our previous works [48,49], showed that the evolution of carbon monoxide occurs concurrently with hydrogen uptake, especially for high alkali loading. This indicates the presence of surface carbonates on the catalyst before H2 activation, as previously noted [52,53]. Doping with K or Rb increases surface basicity, such that more carbonate (i.e., adsorbed carbon dioxide—which is acidic) decomposes from the catalyst as the alkali loading is increased. Decomposition of these surface carbonates through Pt-assisted decarbonylation is easily monitored by in situ DRIFTS (Figure 3) with the generation of CO via Pt carbonyl species.
Table 2 provides assignments from the open literature regarding adsorbed species during ESR. Ethanol adsorption during DRIFTS was found to produce two bands in the range 1000–1200 cm−1 (Figure 4, Figure 5, Figure 6 and Figure 7 for K series and Figure 8, Figure 9, Figure 10 and Figure 11 for Rb series), which are assigned to ethoxy species. These species are formed by the dissociative adsorption of ethanol on the catalyst surface [54,55,56]. Type II ethoxy species located at surface defects on zirconia exhibit a low wavenumber ν(CO) band at ~1050 cm−1, while Type I ethoxy species are associated with unreduced sites on zirconia and positioned at higher wavenumbers [23]. Acetate produced several observable bands: symmetric υ(OCO) stretching (1300 cm−1), asymmetric υ(OCO) stretching (1510 cm−1), and υ(C–H) stretching bands (2700–3100 cm−1). These assignments were confirmed by comparing our results with the literature [2]. The observed formation of acetate likely suggests that ethoxy species underwent oxidative dehydrogenation.
DRIFTS of transient ESR was conducted on unpromoted, K-promoted, and Rb-promoted catalysts to shed further light on the possible mechanism (Figure 4, Figure 5, Figure 6, Figure 7, Figure 8, Figure 9, Figure 10 and Figure 11). Bands of ethoxy species, formed from dissociative adsorption of ethanol, were observed in the range of 50–150 °C for the unpromoted catalyst (Figure 4 and Figure 8 for K-doped and Rb-doped series, respectively). Increasing the temperature further, the surface concentration of ethoxy species decreased until they were virtually completely decomposed, while, simultaneously, there was a concomitant increase in the intensity of bands assigned to acetate; this suggests that ethoxy species underwent oxidative dehydrogenation to acetate—during this stage, CO2 gas was not produced. At the same time, the magnitude of the Pt-CO band increased, suggesting that decarbonylation occurred to a certain extent. After the amount of acetate attained a maximum of approximately 250–300 °C, further steam reforming afforded CH4 and CO2 in addition to CO. The detection of CH4 and CO2 is consistent with the forward decomposition of acetate. The acetate species nearly completely decomposed by 400 °C.
The effect of alkali promotion was also explored by conducting DRIFTS and varying the K loading (Figure 5, Figure 6 and Figure 7) and Rb loading (Figure 9, Figure 10 and Figure 11). DRIFTS revealed that the ethoxy intermediate reacted most rapidly on the 2.55% K- and 4.65% Rb-doped catalyst, evidenced by the fact that it was entirely converted to acetate by 150 °C through oxidative dehydrogenation. In contrast, at low-potassium (0.85%) or -rubidium (0.93%) loading, the ethoxy species exhibited greater stability, as they had significantly decomposed by only 200 °C, very similar to what was observed in the case of the undoped 2% Pt/m-ZrO2 catalyst. The acetate species is observed for all the potassium or rubidium loadings, suggesting it is a likely key intermediate during ESR. Once acetate is formed, however, DRIFTS results showed some differences among catalysts in terms of the selectivity of acetate decomposition, which depended on potassium or rubidium loading. The Pt–carbonyl bands for undoped, 0.85% K, and 0.93% Rb catalysts were at higher wavenumbers and at a significantly increased intensity, reaching maxima at 150–200 °C for the undoped, 150 °C for the 0.85% K doped, and 150–200 °C for the 0.93% Rb catalysts, respectively. This is reasonable to expect since acetate decomposes through different pathways depending on the level of alkali promotion. DRIFTS results suggest that acetate decarboxylation is preferred at high K (2.55% K and 4.25% K) or Rb (4.65% and 9.29% Rb) loading, while the decarbonylation is more favored on the unpromoted catalyst and lower loading alkali-doped catalysts (0.85% K and 0.93% Rb). The dependence of the two different decomposition pathways on alkali loading was also observed for the analogous Na-doped system for both ESR [40,41] and methanol steam reforming [57]. In the latter case, the analogous formate intermediate species were observed (formed from oxidative dehydrogenation of methoxy species), which decarboxylated at 2.5% Na loading, whereas it decarbonylated at low Na loading (i.e., 0.25%, 0.5%, 1%). This decarboxylation pathway enhanced CO2 selectivity and CO conversion, making 1.8% Na-2.5% Na the optimum loading range for H2 production.
DRIFTS results showed that acetate species decomposed at a lower temperature for the 2.55% K-doped catalyst and 4.65% Rb-doped catalyst as compared to the unpromoted catalysts, as a significant band for CH4 was detected at 150 °C (with a slight signal even at 100 °C), while a less intense band for CH4 was observed at 200 °C for the unpromoted catalysts (with a slight signal even detected at 150 °C). Thus, the acetate C–C bond breaks at lower temperatures for the K- and Rb-promoted catalysts.
This bond is analogous to the formate C–H band that is seen during water–gas shift or methanol steam reforming. In our prior WGS investigations, increasing Na, K, Rb, or Cs loading shifted the ν(CH) band to lower wavenumbers. At the same time, the difference between the ν(OCO) bands for asymmetric and symmetric stretching increased with alkali loading. This suggests that changes in surface basicity may be responsible for the CH bond weakening of formate. In the case of ESR, C–C bond weakening is not easily measured. However, we indeed see an increase in the wavenumber difference between the ν(OCO) bands for asymmetric and symmetric stretching. From Table 2, they are: 0% K, ∆ = 117 cm−1; 0.85% K, ∆ = 120 cm−1; 2.55% K, ∆ = 172 cm−1; 4.25% K, ∆ = 172 cm−1; 0% Rb, ∆ = 115 cm−1; 0.93% Rb, ∆ = 119 cm−1; 4.65% Rb, ∆ = 175 cm−1; and 9.29% Rb, ∆ = 176 cm−1. In all cases, the results suggest that the alkali promotes the weakening of the respective bond, emphasizing once again the analogous nature of the two alkali-doped catalyst systems. At 4.25% K or 9.29% Rb doping level, the catalyst surface was significantly blocked with the alkali, thereby creating a bottleneck for the formation of, and subsequent decomposition of, the intermediates. In addition to the CO2 that evolved, residual carbonates are clearly observed on the catalysts with K or Rb, as bands such as at 1620, 1574–1439 cm−1, and ~1315 cm−1 O–C–O stretching modes of carbonates [58].
Interestingly, the optimum K loading (2.55%) corresponds to a very similar weight percent as the optimal Na loading (2.5%) from prior work [41], meaning that the optimal K loading occurred at 60% of the optimal Na loading atomically. Some works have shown that at excessive alkali loadings on Pt/ZrO2, the surface of the Pt nanoparticles is blocked, inhibiting the role of Pt in hydrogen transfer reactions during LT-WGS [59,60]. This effect provides a reasonable explanation for the results obtained, given that K+ is considerably larger than Na+; the surface of the Pt nanoparticles becomes covered by the alkali metal at considerably lower atomic loadings for potassium-promoted catalysts. Additionally, it has been shown that at high alkali loadings, catalyst basicity promotes the formation of the carbonate intermediate, which is the precursor to CO2 formation; however, it was also found to impede CO2 liberation during LT-WGS [60,61]. This may be due to higher basicity since CO2 is an acidic molecule that is therefore adsorbed more strongly and/or it may be due to the fact that the alkali obstructs the metallic function, which is known to facilitate carbonate decomposition. Due to the fact that potassium is a more basic promoter than sodium, one would expect this effect to cause CO2 liberation to be significantly hindered for potassium relative to sodium. However, potassium is able to achieve similar promotion to sodium, and it does so at a lower atomic loading; this may be due to its lower electronegativity.
DRIFTS experiments showed that the optimal loading for the Rb-promoted catalyst is 4.65%, which is 80% and 50% of the optimal K and Na atomically loading, respectively. This trend suggests that a lower alkali atomically loading is required for the C–C bond scission of intermediate acetate when the alkali electronegativity decreases. Indeed, the electronegativity of Rb is 0.706 compared to 0.869 for Na and 0.734 for K [62]. A similar trend also was observed in our previous work [48]. Decreasing the electronegativity, the ν(C–H) formate band progressively shifted to lower wavenumbers, which was associated with a faster formate decomposition in the presence of steam [59,63,64].
Figure 12 and Figure 13 show the TPD-MS profiles for methane over the K and Rb series of catalysts, respectively. The main peak for the 2% Pt/ZrO2 catalyst is at 391 °C, with a very minor peak at 200 °C. The position of the main peak did not change to a significant degree for the 0.85% K- and 1.70% K-doped catalysts, with peaks at 383 °C and 360 °C, respectively; minor peaks at low temperature were also observed at 100–190 °C and 115–170 °C. However, a significant shift to lower temperature occurred once the dopant loading reached 2.55% K, where the temperature was 270 °C for the main peak, with a low-temperature peak at 130 °C. At 3.40% K, 4.25% K, and 8.50% K doping levels, the main peak decreased in intensity and shifted to higher temperatures of 314 °C, 327 °C, and 345 °C, respectively, while the low-temperature peak remained at a similar temperature. The results indicate that cleaving of the C–C bond of acetate species during ESR is not only promoted by alkali addition but also that the K-doping loading is close to the optimum at 2.55% K. For the Rb series, increasing Rb loading to 1.86% Rb decreases the main CH4 evolution signal from 391 °C to 384 °C. In the range of 2.79% Rb–4.65% Rb, the peak at 370 °C is attenuated, and a new low-temperature peak at 160–180 °C emerges. At 5.58% and 9.29% Rb loadings, the CH4 signal is attenuated overall, and by 9.3% Rb, the signal of the higher temperature peak has increased to 430 °C due to blocking of the catalyst surface by excessive alkali. DRIFTS and TPD-MS show that alkali promotion had a beneficial effect on the C–C bond scission of the intermediate acetate. However, the electronic effect of alkali on the structure of Pt metal is not well understood. Different hypotheses can be formulated: (a) charge transfer, which may result in a change in the white line intensity in the presence of K [65]; (b) an electrostatic effect, which could cause bond weakening in adsorbed species on the catalyst surface [66]; (c) a Fermi level electronic perturbation [43]; or (d) an alteration in bond strengths due to changes in the acidity/basicity of the catalyst. Confounding the analysis of (a) is particle size effects [67,68], which tend to cause the binding energy and white line to both increase with decreasing particle size. To gain insight into whether K or Rb promotion leads to charge transfer from the alkali to Pt, the Pt L-3 XANES spectra can be analyzed. As shown in Figure 14 for the K-series and Figure 15 for the Rb-series, during reduction in hydrogen at 350 °C and after cooling in hydrogen to ambient conditions, it is evident that there are differences in the XANES spectra of catalysts having no or low alkali content versus those with high alkali content. However, the white line intensity is also affected by the size of Pt particles, so the difference between the L-3 XANES and L-2 XANES can be studied to remove this effect (Figure 14 and Figure 15). If K or Rb promotion causes electron charge transfer from the alkali to Pt, then the Pt L3-L2 XANES difference should decrease in magnitude as a function of alkali loading. However, this trend is not observed, indicating that neither potassium nor rubidium is likely transferring electron charge density to platinum. Despite this, it is still suggested that the alkali and Pt are in direct contact, which can be seen through the TPR-XANES and TPR-EXAFS reported in our previous studies [48,49]. The TPR-XANES and TPR-EXAFS revealed that as the loading of K or Rb increases, the reduction of PtO is hindered, which is likely the result of the covering of platinum by the alkali. It is also expected that if the alkali donated electron density to Pt particles, a relaxation in the edge energy should occur (related to the binding energy) [69,70]. Figure 16 and Figure 17 reveal that in comparing catalysts with the Pt0 foil, no such shift was detected.
The catalytic activity data for K and Rb-series are reported in Table 3, whereas the H2 yield and product selectivity trends of the unpromoted catalyst at different conversions are reported in Table 4. The addition of potassium or rubidium progressively decreased the ethanol conversion. A similar trend was observed for sodium in our previous work [41]. Catalysts with very high alkali loading (i.e., 4.25% K or 9.29% Rb) exhibited negligible catalytic activity. TEM-EDX and EXAFS fittings showed that Pt clusters aggregated at higher loading. Moreover, the alkali likely partially covers the platinum particles, as evidenced by a decreasing ν(CO) band at higher K or Rb alkali loading (virtually disappearing at the highest loadings). However, the most interesting effect of alkali promotion on Pt/ZrO2 is related to product selectivity. Indeed, CO is only detected among the products for unpromoted and low alkali doping (0.85% K and 0.93% Rb), whereas no CO is produced at higher alkali loading. Alkali promotion decreased the acetaldehyde selectivity, which is 3.49% for the unpromoted, whereas it is lower than 1.5% for the K- or Rb-promoted catalyst. The activity results confirm that different pathways occur depending on the potassium or rubidium loading, as already pointed out by DRIFTS. The decarbonylation route is present for the unpromoted catalyst and the catalysts with low alkali loading, whereas the decarboxylation route (where acetate decomposes to CH4 and a carbonate species, which further decomposes to CO2) completely dominates when the alkali loadings reach 2.55% K and 4.93% Rb; a similar effect occurred at 1.80% Na in our prior work [41]. Moreover, 2.55% K corresponds to 5.57% Rb and 1.50% Na on an atomic loading basis. Therefore, the results suggest that decarboxylation is improved by increasing the basicity of the alkali (moving down the Group 1 column) because the atomic loading of potassium to stave off decarbonylation is lower than that of sodium and higher than that of rubidium.
However, Table 4 shows that the product selectivity of Pt/ZrO2 changes to a degree as a function of conversion. Therefore, in order to place alkali effects on a firmer footing, it is necessary to compare the product selectivity and H2 yield at the same conversion level. As shown in Table 5, the alkali-promoted catalysts have approximately 60–62% higher H2 yield as compared to the unpromoted catalyst when compared at similar ethanol conversion. This is due in part to enhanced decarboxylation over decarbonylation, as the CO2 selectivity is increased by 240–273%, while the CO selectivity is diminished by 45–61%. While the 0.85% K and 0.93% Rb catalysts provide similar H2 yields, 0.85% K corresponds to 1.86% Rb. The fact that a similar improvement in H2 yield occurred over that of the unpromoted 2% Pt/ZrO2 catalyst at an atomic loading of Rb that is 50% of the atomic loading of K indicates that higher basicity alkali promoters are more effective at facilitating the more selective decarboxylation pathway.

4. Conclusions

The addition of potassium or rubidium to Pt/ZrO2 progressively decreased the surface area and the pore volume because of some pore blocking. Platinum particle size was ~1 nm for the unpromoted and lower alkali loading (0.85% K and 0.93% Rb), while aggregation occurred at higher alkali loading. TEM-EDX of 9.29% Rb-Pt/ZrO2 showed Pt aggregates of 10 nm. The difference between Pt L3–L2 XANES spectra indicates that neither potassium nor rubidium is likely transferring electron charge density to platinum. Moreover, no relaxation effect on the edge jump energy was observed with the addition of K or Rb.
DRIFTS experiments were carried out to investigate the mechanism. The results suggest that the catalysts have similar steps during ESR: dissociation of ethanol to produce an ethoxy species, oxidative dehydrogenation of ethoxy species to acetate, and acetate decomposition. The forward decomposition of acetate to CH4 and carbonate (the precursor to CO2) is facilitated by the presence of K or Rb, and there is an optimum alkali loading for facilitating the C–C scission of acetate. This is inferred from the temperature at which CH4 evolution occurs as well as a systematic increase in the difference in band position for ν (OCO) asymmetric and symmetric stretching for acetate, which occurs with increasing the alkali loading. Ethoxy species are more stable on the unpromoted catalyst and acetate decomposition, which is associated with the formation of methane, occurs at a higher temperature. Methane formation was detected at 391 °C for 2% Pt/ZrO2 in TPD-MS, whereas it occurred at 270 °C for 2.55% K and in multiple peaks (160–180 °C and 370 °C) for 2.79–4.65% Rb. Moreover, DRIFTS experiments and catalytic activity testing point out the existence of different pathways for acetate decomposition depending on the alkali loading. Decarboxylation is the most favored route at high alkali loading (2.55 and 4.25 wt.%). In this pathway, acetate decomposes in the forward direction, yielding CH4 and a carbonate species, which further decomposes to CO2. In contrast, the unselective decarbonylation pathway occurs to a significant extent for the unpromoted catalyst and the catalyst having low alkali loading (e.g., 0.85% K or 0.93% Rb). By increasing the alkali basicity by switching from K to Rb, lower loadings of alkali enabled virtually complete blocking of the non-selective decarbonylation pathway. Results of unpromoted, K-promoted, and Rb-promoted catalysts compared at similar conversion further confirmed the promoting effect of the alkali, as well as the basicity trend.

Author Contributions

Conceptualization, catalyst preparation, catalyst characterization, formal analysis, supervision, and writing, G.J. Catalyst preparation, catalyst characterization, formal analysis, and writing, R.G. and C.D.W. Reaction testing, characterization, formal analysis, conceptualization, and writing, M.M. Catalyst preparation, supervision, and resources, D.C.C. Catalyst characterization, data curation, resources, and supervision, A.J.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the USDA National Institute of Food and Agriculture, Interdisciplinary Hands-on Research Traineeship and Extension Experiential Learning in Bioenergy/Natural Resources/Economics/Rural project, U-GREAT (Under Graduate Research, Education And Training) program (2016-67032-24984). This research was also supported by the National Science Foundation through Grant Award 1832388.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Argonne’s research was supported in part by the US Department of Energy (DOE), Office of Fossil Energy, National Energy Technology Laboratory (NETL). Advanced photon source was supported by the US Department of Energy, Office of Science, Office of Basic Energy Sciences, under contract number DE-AC02-06CH11357. MRCAT operations are supported by the Department of Energy and the MRCAT member institutions. CAER research was supported by the Commonwealth of Kentucky. Research carried out at UTSA was supported by UTSA, the State of Texas, and the STARs program. Caleb D. Watson would like to acknowledge funding from the UTSA College of Engineering Scholarship. Gary Jacobs would like to thank UTSA and the State of Texas for financial support through startup funds.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, D.; Li, X.; Gong, J. Catalytic Reforming of Oxygenates: State of the Art and Future Prospects. Chem. Rev. 2016, 116, 11529–11653. [Google Scholar] [CrossRef] [Green Version]
  2. Mattos, L.V.; Jacobs, G.; Davis, B.H.; Noronha, F.B. Production of Hydrogen from Ethanol: Review of Reaction Mechanism and Catalyst Deactivation. Chem. Rev. 2012, 112, 4094–4123. [Google Scholar] [CrossRef] [PubMed]
  3. Ni, M.; Leung, D.Y.C.; Leung, M.K.H. A review on reforming bio-ethanol for hydrogen production. Int. J. Hydrogen Energy 2007, 32, 3238–3247. [Google Scholar] [CrossRef]
  4. Vaidya, P.D.; Rodrigues, A.E. Glycerol Reforming for Hydrogen Production: A Review. Chem. Eng. Tech. 2009, 32, 1463–1469. [Google Scholar] [CrossRef]
  5. Tran, N.H.; Kannangara, G.S.K. Conversion of glycerol to hydrogen rich gas. Chem. Soc. Rev. 2013, 42, 9454–9479. [Google Scholar] [CrossRef]
  6. Słowik, G.; Greluk, M.; Rotko, M.; Machocki, A. Evolution of the structure of unpromoted and potassium-promoted ceria-supported nickel catalysts in the steam reforming of ethanol. Appl. Catal. B Environ. 2018, 221, 490–509. [Google Scholar] [CrossRef]
  7. Contreras, J.L.; Salmones, J.; Colín-Luna, J.A.; Nuño, L.; Quintana, B.; Córdova, I.; Zeifert, B.; Tapia, C.; Fuentes, G.A. Catalysts for H2 production using the ethanol steam reforming (a review). Int. J. Hydrogen Energy 2014, 39, 18835–18853. [Google Scholar] [CrossRef]
  8. Ogo, S.; Sekine, Y. Recent progress in ethanol steam reforming using non-noble transition metal catalysts: A review. Fuel Proc. Tech. 2020, 199, 106238. [Google Scholar] [CrossRef]
  9. Frusteri, F.; Freni, S.; Spadaro, L.; Chiodo, V.; Bonura, G.; Donato, S.; Cavallaro, S. H2 production for MC fuel cell by steam reforming of ethanol over MgO supported Pd, Rh, Ni and Co catalysts. Catal. Commun. 2004, 5, 611–615. [Google Scholar] [CrossRef]
  10. Song, H.; Ozkan, U.S. Ethanol steam reforming over Co-based catalysts: Role of oxygen mobility. J. Catal. 2009, 261, 66–74. [Google Scholar] [CrossRef]
  11. Ferencz, Z.; Varga, E.; Puskás, R.; Kónya, Z.; Baán, K.; Oszkó, A.; Erdőhelyi, A. Reforming of ethanol on Co/Al2O3 catalysts reduced at different temperatures. J. Catal. 2018, 358, 118–130. [Google Scholar] [CrossRef]
  12. Gaudillere, C.; González, J.J.; Chica, A.; Serra, J.M. YSZ monoliths promoted with Co as catalysts for the production of H2 by steam reforming of ethanol. Appl. Catal. A Gen. 2017, 538, 165–173. [Google Scholar] [CrossRef]
  13. Ángel-Soto, J.; Martínez-Rosales, M.; Ángel-Soto, P.; Zamorategui-Molina, A. Synthesis, characterization and catalytic application of Ni catalysts supported on alumina–zirconia mixed oxides. Bullet. Mat. Sci. 2017, 40, 1309–1318. [Google Scholar] [CrossRef] [Green Version]
  14. Campos, C.H.; Pecchi, G.; Fierro, J.L.G.; Osorio-Vargas, P. Enhanced bimetallic Rh-Ni supported catalysts on alumina doped with mixed lanthanum-cerium oxides for ethanol steam reforming. Molec. Catal. 2019, 469, 87–97. [Google Scholar] [CrossRef]
  15. Liguras, D.K.; Kondarides, D.I.; Verykios, X.E. Production of hydrogen for fuel cells by steam reforming of ethanol over supported noble metal catalysts. Appl. Catal. B Environ. 2003, 43, 345–354. [Google Scholar] [CrossRef]
  16. Yamazaki, T.; Kikuchi, N.; Katoh, M.; Hirose, T.; Saito, H.; Yoshikawa, T.; Wada, M. Behavior of steam reforming reaction for bio-ethanol over Pt/ZrO2 catalysts. Appl. Catal. B Environ. 2010, 99, 81–88. [Google Scholar] [CrossRef]
  17. De Lima, S.M.; Silva, A.M.; da Cruz, I.O.; Jacobs, G.; Davis, B.H.; Mattos, L.V.; Noronha, F.B. H2 production through steam reforming of ethanol over Pt/ZrO2, Pt/CeO2 and Pt/CeZrO2 catalysts. Catal. Today 2008, 138, 162–168. [Google Scholar] [CrossRef]
  18. Bilal, M.; Jackson, S.D. Ethanol steam reforming over Rh and Pt catalysts: Effect of temperature and catalyst deactivation. Catal. Sci. Technol. 2013, 3, 754–766. [Google Scholar] [CrossRef]
  19. De Lima, S.M.; Silva, A.M.; Graham, U.M.; Jacobs, G.; Davis, B.H.; Mattos, L.V.; Noronha, F.B. Ethanol decomposition and steam reforming of ethanol over CeZrO2 and Pt/CeZrO2 catalyst: Reaction mechanism and deactivation. Appl. Catal. A Gen. 2009, 352, 95–113. [Google Scholar] [CrossRef]
  20. Jacobs, G.; Davis, B.H. In situ DRIFTS investigation of the steam reforming of methanol over Pt/ceria. Appl. Catal. A Gen. 2005, 285, 43–49. [Google Scholar] [CrossRef]
  21. Ciambelli, P.; Palma, V.; Ruggiero, A. Low temperature catalytic steam reforming of ethanol. 1. The effect of the support on the activity and stability of Pt catalysts. Appl. Catal. B Environ. 2010, 96, 18–27. [Google Scholar] [CrossRef]
  22. He, Z.; Yang, M.; Wang, X.; Zhao, Z.; Duan, A. Effect of the transition metal oxide supports on hydrogen production from bio-ethanol reforming. Catal. Today 2012, 194, 2–8. [Google Scholar] [CrossRef]
  23. Jacobs, G.; Keogh, R.A.; Davis, B.H. Steam reforming of ethanol over Pt/ceria with co-fed hydrogen. J. Catal. 2007, 245, 326–337. [Google Scholar] [CrossRef]
  24. Shido, T.; Iwasawa, Y. Reactant-Promoted Reaction Mechanism for Water-Gas Shift Reaction on Rh-Doped CeO2. J. Catal. 1993, 141, 71–81. [Google Scholar] [CrossRef]
  25. Jacobs, G.; Graham, U.M.; Chenu, E.; Patterson, P.M.; Dozier, A.; Davis, B.H. Low-temperature water–gas shift: Impact of Pt promoter loading on the partial reduction of ceria and consequences for catalyst design. J. Catal. 2005, 229, 499–512. [Google Scholar] [CrossRef]
  26. Laachir, A.; Perrichon, V.; Badri, A.; Lamotte, J.; Catherine, E.; Lavalley, J.C.; El Fallah, J.; Hilaire, L.; Le Normand, F.; Quéméré, E.; et al. Reduction of CeO2 by hydrogen. Magnetic susceptibility and Fourier-transform infrared, ultraviolet and X-ray photoelectron spectroscopy measurements. J. Chem. Soc. Faraday Trans. 1991, 87, 1601–1609. [Google Scholar] [CrossRef]
  27. Grzybek, G.; Greluk, M.; Indyka, P.; Góra-Marek, K.; Legutko, P.; Słowik, G.; Turczyniak-Surdacka, S.; Rotko, M.; Sojka, Z.; Kotarba, A. Cobalt catalyst for steam reforming of ethanol–Insights into the promotional role of potassium. Int. J. Hydrogen Energy 2020, 45, 22658–22673. [Google Scholar] [CrossRef]
  28. Grzybek, G.; Góra-Marek, K.; Patulski, P.; Greluk, M.; Rotko, M.; Słowik, G.; Kotarba, A. Optimization of the potassium promotion of the Co|α-Al2O3 catalyst for the effective hydrogen production via ethanol steam reforming. Appl. Catal. A Gen. 2021, 614, 118051. [Google Scholar] [CrossRef]
  29. Llorca, J.; Homs, N.S.; Sales, J.; Fierro, J.-L.G.; Ramírez de la Piscina, P. Effect of sodium addition on the performance of Co–ZnO-based catalysts for hydrogen production from bioethanol. J. Catal. 2004, 222, 470–480. [Google Scholar] [CrossRef]
  30. Espinal, R.; Taboada, E.; Molins, E.; Chimentao, R.J.; Medina, F.; Llorca, J. Cobalt hydrotalcites as catalysts for bioethanol steam reforming. The promoting effect of potassium on catalyst activity and long-term stability. Appl. Catal. B Environ. 2012, 127, 59–67. [Google Scholar] [CrossRef]
  31. Ogo, S.; Shimizu, T.; Nakazawa, Y.; Mukawa, K.; Mukai, D.; Sekine, Y. Steam reforming of ethanol over K promoted Co catalyst. Appl. Catal. A Gen. 2015, 495, 30–38. [Google Scholar] [CrossRef]
  32. Yoo, S.; Park, S.; Song, J.H.; Kim, D.H. Hydrogen production by the steam reforming of ethanol over K-promoted Co/Al2O3–CaO xerogel catalysts. Molec. Catal. 2020, 491, 110980. [Google Scholar] [CrossRef]
  33. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; Di Blasi, O.; Bonura, G.; Cavallaro, S. Steam reforming of bio-ethanol on alkali-doped Ni/MgO catalysts: Hydrogen production for MC fuel cell. Appl. Catal. A Gen. 2004, 270, 1–7. [Google Scholar] [CrossRef]
  34. Dömök, M.; Baán, K.; Kecskés, T.; Erdőhelyi, A. Promoting Mechanism of Potassium in the Reforming of Ethanol on Pt/Al2O3 Catalyst. Catal. Lett. 2008, 126, 49–57. [Google Scholar] [CrossRef]
  35. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; Bonura, G.; Cavallaro, S. Potassium improved stability of Ni/MgO in the steam reforming of ethanol for the production of hydrogen for MCFC. J. Power Sources 2004, 132, 139–144. [Google Scholar] [CrossRef]
  36. Sharma, Y.C.; Kumar, A.; Prasad, R.; Upadhyay, S.N. Ethanol steam reforming for hydrogen production: Latest and effective catalyst modification strategies to minimize carbonaceous deactivation. Renew. Sustain. Energy Rev. 2017, 74, 89–103. [Google Scholar] [CrossRef]
  37. Greluk, M.; Rybak, P.; Słowik, G.; Rotko, M.; Machocki, A. Comparative study on steam and oxidative steam reforming of ethanol over 2KCo/ZrO2 catalyst. Catal. Today 2015, 242, 50–59. [Google Scholar] [CrossRef]
  38. Banach, B.; Machocki, A. Effect of potassium addition on a long term performance of Co–ZnO–Al2O3 catalysts in the low-temperature steam reforming of ethanol: Co-precipitation vs citrate method of catalysts synthesis. Appl. Catal. A Gen. 2015, 505, 173–182. [Google Scholar] [CrossRef]
  39. Słowik, G.; Gawryszuk-Rżysko, A.; Greluk, M.; Machocki, A. Estimation of Average Crystallites Size of Active Phase in Ceria-Supported Cobalt-Based Catalysts by Hydrogen Chemisorption vs TEM and XRD Methods. Catal. Lett. 2016, 146, 2173–2184. [Google Scholar] [CrossRef] [Green Version]
  40. Martinelli, M.; Watson, C.D.; Jacobs, G. Sodium doping of Pt/m-ZrO2 promotes C–C scission and decarboxylation during ethanol steam reforming. Int. J. Hydrogen Energy 2020, 45, 18490–18501. [Google Scholar] [CrossRef]
  41. Martinelli, M.; Castro, J.D.; Alhraki, N.; Matamoros, M.E.; Kropf, A.J.; Cronauer, D.C.; Jacobs, G. Effect of sodium loading on Pt/ZrO2 during ethanol steam reforming. Appl. Catal. A Gen. 2021, 610, 117947. [Google Scholar] [CrossRef]
  42. Roh, H.-S.; Platon, A.; Wang, Y.; King, D.L. Catalyst deactivation and regeneration in low temperature ethanol steam reforming with Rh/CeO2–ZrO2 catalysts. Catal. Lett. 2006, 110, 1–6. [Google Scholar] [CrossRef]
  43. Ramaker, D.E.; Mojet, B.L.; Garriga Oostenbrink, M.T.; Miller, J.T.; Koningsberger, D.C. Contribution of shape resonance and Pt–H EXAFS in the Pt L2,3 X-ray absorption edges of supported Pt particles: Application and consequences for catalyst characterization. Phys. Chem. Chem. Phys. 1999, 1, 2293–2302. [Google Scholar] [CrossRef]
  44. Ressler, T. WinXAS: A Program for X-ray Absorption Spectroscopy Data Analysis under MS-Windows. J. Synchrotron Rad. 1998, 5, 118–122. [Google Scholar] [CrossRef] [PubMed]
  45. Ravel, B. ATOMS: Crystallography for the X-ray absorption spectroscopist. J. Synchrotron Radiat. 2001, 8, 314–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Newville, M.; Ravel, B.; Haskel, D.; Rehr, J.J.; Stern, E.A.; Yacoby, Y. Analysis of multiple-scattering XAFS data using theoretical standards. Phys. B Cond. Matt. 1995, 208–209, 154–156. [Google Scholar] [CrossRef]
  47. Martinelli, M.; Jacobs, G.; Graham, U.M.; Shafer, W.D.; Cronauer, D.C.; Kropf, A.J.; Marshall, C.L.; Khalid, S.; Visconti, C.G.; Lietti, L.; et al. Water-gas shift: Characterization and testing of nanoscale YSZ supported Pt catalysts. Appl. Catal. A Gen. 2015, 497, 184–197. [Google Scholar] [CrossRef] [Green Version]
  48. Watson, C.D.; Martinelli, M.; Cronauer, D.C.; Kropf, A.J.; Marshall, C.L.; Jacobs, G. Low temperature water-gas shift: Optimization of K loading on Pt/m-ZrO2 for enhancing CO conversion. Appl. Catal. A Gen. 2020, 598, 117572. [Google Scholar] [CrossRef]
  49. Watson, C.D.; Martinelli, M.; Cronauer, D.C.; Kropf, A.J.; Jacobs, G. Low Temperature Water-Gas Shift: Enhancing Stability through Optimizing Rb Loading on Pt/ZrO2. Catalysts 2021, 11, 210. [Google Scholar] [CrossRef]
  50. Jentys, A. Estimation of mean size and shape of small metal particles by EXAFS. Phys. Chem. Chem. Phys. 1999, 1, 4059–4063. [Google Scholar] [CrossRef]
  51. Marinkovic, N.S.; Sasaki, K.; Azic, R.R. Nanoparticle size evaluation of catalysts by EXAFS: Advantages and limitations. Zast. Mater. 2016, 57, 101–109. [Google Scholar] [CrossRef] [Green Version]
  52. Chenu, E.; Jacobs, G.; Crawford, A.C.; Keogh, R.A.; Patterson, P.M.; Sparks, D.E.; Davis, B.H. Water-gas shift: An examination of Pt promoted MgO and tetragonal and monoclinic ZrO2 by in situ DRIFTS. Appl. Catal. B Environ. 2005, 59, 45–56. [Google Scholar] [CrossRef]
  53. Pigos, J.M.; Brooks, C.J.; Jacobs, G.; Davis, B.H. Low temperature water-gas shift: Characterization of Pt-based ZrO2 catalyst promoted with Na discovered by combinatorial methods. Appl. Catal. A Gen. 2007, 319, 47–57. [Google Scholar] [CrossRef]
  54. Yee, A.; Morrison, S.J.; Idriss, H. A Study of Ethanol Reactions over Pt/CeO2 by Temperature-Programmed Desorption and in Situ FT-IR Spectroscopy: Evidence of Benzene Formation. J. Catal. 2000, 191, 30–45. [Google Scholar] [CrossRef]
  55. Yee, A.; Morrison, S.J.; Idriss, H. A Study of the Reactions of Ethanol on CeO2 and Pd/CeO2 by Steady State Reactions, Temperature Programmed Desorption, and In Situ FT-IR. J. Catal. 1999, 186, 279–295. [Google Scholar] [CrossRef]
  56. Mattos, L.V.; Noronha, F.B. Hydrogen production for fuel cell applications by ethanol partial oxidation on Pt/CeO2 catalysts: The effect of the reaction conditions and reaction mechanism. J. Catal. 2005, 233, 453–463. [Google Scholar] [CrossRef]
  57. Martinelli, M.; Jacobs, G.; Graham, U.M.; Davis, B.H. Methanol Steam Reforming: Na Doping of Pt/YSZ Provides Fine Tuning of Selectivity. Catalysts 2017, 7, 148. [Google Scholar] [CrossRef] [Green Version]
  58. Binet, C.; Daturi, M.; Lavalley, J.-C. IR study of polycrystalline ceria properties in oxidised and reduced states. Catal. Today 1999, 50, 207–225. [Google Scholar] [CrossRef]
  59. Martinelli, M.; Alhraki, N.; Castro, J.D.; Matamoros, M.E.; Jacobs, G. New Dimensions in Production and Utilization of Hydrogen, 1st ed.; Nanda, S., Vo, D.V., Nguyen-Tri, P., Eds.; Elsevier International Publishing: Amsterdam, The Netherlands, 2020; pp. 143–160. [Google Scholar]
  60. Evin, H.N.; Jacobs, G.; Ruiz-Martinez, J.; Graham, U.M.; Dozier, A.; Thomas, G.; Davis, B.H. Low Temperature Water–Gas Shift/Methanol Steam Reforming: Alkali Doping to Facilitate the Scission of Formate and Methoxy C–H Bonds over Pt/ceria Catalyst. Catal. Lett. 2008, 122, 9–19. [Google Scholar] [CrossRef]
  61. Evin, H.N.; Jacobs, G.; Ruiz-Martinez, J.; Thomas, G.A.; Davis, B.H. Low Temperature Water–Gas Shift: Alkali Doping to Facilitate Formate C–H Bond Cleaving over Pt/Ceria Catalysts—An Optimization Problem. Catal. Lett. 2008, 120, 166–178. [Google Scholar] [CrossRef]
  62. Allen, L.C. Electronegativity is the average one-electron energy of the valence-shell electrons in ground-state free atoms. J. Amer. Chem. Soc. 1989, 111, 9003–9014. [Google Scholar] [CrossRef]
  63. Pigos, J.M.; Brooks, C.J.; Jacobs, G.; Davis, B.H. Low temperature water–gas shift: The effect of alkali doping on the CH bond of formate over Pt/ZrO2 catalysts. Appl. Catal. A Gen. 2007, 328, 14–26. [Google Scholar] [CrossRef]
  64. Martinelli, M.; Jacobs, G.; Shafer, W.D.; Davis, B.H. Effect of alkali on CH bond scission over Pt/YSZ catalyst during water-gas-shift, steam-assisted formic acid decomposition and methanol steam reforming. Catal. Today 2017, 291, 29–35. [Google Scholar] [CrossRef]
  65. Menacherry, P.V.; Haller, G.L. Electronic effects and effects of particle morphology in n-hexane conversion over zeolite-supported platinum catalysts. J. Catal. 1998, 177, 175–188. [Google Scholar] [CrossRef]
  66. Fukunaga, T.; Ponec, V. On the role of additives to platinum catalysts for reforming reactions. Appl. Catal. A Gen. 1997, 154, 207–219. [Google Scholar] [CrossRef]
  67. Bazin, D.; Sayers, D.; Rehr, J.J.; Mottet, C. Numerical Simulation of the Platinum LIII Edge White Line Relative to Nanometer Scale Clusters. J. Phys. Chem. B 1997, 101, 5332–5336. [Google Scholar] [CrossRef]
  68. Dai, Y.; Gorey, T.J.; Anderson, S.L.; Lee, S.; Lee, S.; Seifert, S.; Winans, R.E. Inherent Size Effects on XANES of Nanometer Metal Clusters: Size-Selected Platinum Clusters on Silica. J. Phys. Chem. C 2017, 121, 361–374. [Google Scholar] [CrossRef]
  69. Haller, G.L. New catalytic concepts from new materials: Understanding catalysis from a fundamental perspective, past, present, and future. J. Catal. 2003, 216, 12–22. [Google Scholar] [CrossRef]
  70. Mojet, B.L.; Miller, J.T.; Ramaker, D.E.; Koningsberger, D.C. A new model describing the metal-support interaction in noble metal catalysts. J. Catal. 1999, 186, 373–386. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Transmission electron microscopy (TEM) and scanning transmission microscopy with energy-dispersive X-ray spectroscopy (STEM-EDX) images for the (a) 0.93% Rb-2% Pt/ZrO2 catalyst and (b) the 9.29% Rb-2% Pt/ZrO2 catalyst.
Figure 1. Transmission electron microscopy (TEM) and scanning transmission microscopy with energy-dispersive X-ray spectroscopy (STEM-EDX) images for the (a) 0.93% Rb-2% Pt/ZrO2 catalyst and (b) the 9.29% Rb-2% Pt/ZrO2 catalyst.
Nanomaterials 11 02233 g001
Figure 2. (a): TPR profiles for K-promoted catalysts: (a) 2% Pt/ZrO2, (b) 0.85% K-2% Pt/ZrO2, (c) 1.70% K-2% Pt/ZrO2, (d) 2.55% K-2% Pt/ZrO2, (e) 3.40% K-2% Pt/ZrO2, (f) 4.25% K-2% Pt/ZrO2, and (g) 8.50% K-2% Pt/ZrO2; (b): TPR profile for Rb-promoted catalysts: (h) 0.55% Rb-2% Pt/ZrO2, (i) 1.86% Rb-2% Pt/ZrO2, (j) 2.79% Rb-2% Pt/ZrO2, (k) 4.65% Rb-2% Pt/ZrO2, and (l) 9.29% Rb-2% Pt/ZrO2. Figure 2a reprinted from [48] with permission from Elsevier, copyright 2020. Figure 2b reprinted from [48] with permission from MDPI, copyright 2021.
Figure 2. (a): TPR profiles for K-promoted catalysts: (a) 2% Pt/ZrO2, (b) 0.85% K-2% Pt/ZrO2, (c) 1.70% K-2% Pt/ZrO2, (d) 2.55% K-2% Pt/ZrO2, (e) 3.40% K-2% Pt/ZrO2, (f) 4.25% K-2% Pt/ZrO2, and (g) 8.50% K-2% Pt/ZrO2; (b): TPR profile for Rb-promoted catalysts: (h) 0.55% Rb-2% Pt/ZrO2, (i) 1.86% Rb-2% Pt/ZrO2, (j) 2.79% Rb-2% Pt/ZrO2, (k) 4.65% Rb-2% Pt/ZrO2, and (l) 9.29% Rb-2% Pt/ZrO2. Figure 2a reprinted from [48] with permission from Elsevier, copyright 2020. Figure 2b reprinted from [48] with permission from MDPI, copyright 2021.
Nanomaterials 11 02233 g002
Figure 3. DRIFTS spectra of carbonate decomposition during activation for (a) 2% Pt/ZrO2, and the same doped with (b) 0.93% Rb, (c) 4.65% Rb, (d) 9.29% Rb, (e) 0.85% K, (f) 2.55% K, (g) 4.25% K, (h) 8.50% K. Figure 3a–d reprinted from [49] with permission from MDPI, copyright 2021. Figure 3e,f reprinted from [48] with permission from Elsevier, copyright 2020.
Figure 3. DRIFTS spectra of carbonate decomposition during activation for (a) 2% Pt/ZrO2, and the same doped with (b) 0.93% Rb, (c) 4.65% Rb, (d) 9.29% Rb, (e) 0.85% K, (f) 2.55% K, (g) 4.25% K, (h) 8.50% K. Figure 3a–d reprinted from [49] with permission from MDPI, copyright 2021. Figure 3e,f reprinted from [48] with permission from Elsevier, copyright 2020.
Nanomaterials 11 02233 g003
Figure 4. DRIFTS spectra of transient ESR over 2% Pt/m-ZrO2 (K-series).
Figure 4. DRIFTS spectra of transient ESR over 2% Pt/m-ZrO2 (K-series).
Nanomaterials 11 02233 g004
Figure 5. DRIFTS spectra of transient ESR over 0.85% K-2% Pt/m-ZrO2.
Figure 5. DRIFTS spectra of transient ESR over 0.85% K-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g005
Figure 6. DRIFTS spectra of transient ESR over 2.55% K-2% Pt/m-ZrO2.
Figure 6. DRIFTS spectra of transient ESR over 2.55% K-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g006
Figure 7. DRIFTS spectra of transient ESR over 4.25% K-2% Pt/m-ZrO2.
Figure 7. DRIFTS spectra of transient ESR over 4.25% K-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g007
Figure 8. DRIFTS spectra of transient ESR over 2% Pt/m-ZrO2 (Rb-series).
Figure 8. DRIFTS spectra of transient ESR over 2% Pt/m-ZrO2 (Rb-series).
Nanomaterials 11 02233 g008
Figure 9. DRIFTS spectra of transient ESR over 0.93% Rb-2% Pt/m-ZrO2.
Figure 9. DRIFTS spectra of transient ESR over 0.93% Rb-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g009
Figure 10. DRIFTS spectra of transient ESR over 4.65% Rb-2% Pt/m-ZrO2.
Figure 10. DRIFTS spectra of transient ESR over 4.65% Rb-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g010
Figure 11. DRIFTS spectra of transient ESR over 9.29% Rb-2% Pt/m-ZrO2.
Figure 11. DRIFTS spectra of transient ESR over 9.29% Rb-2% Pt/m-ZrO2.
Nanomaterials 11 02233 g011
Figure 12. TPD-MS of ethanol steam reforming over (a) 2% Pt/ZrO2, (b) 0.85% K-2% Pt/ZrO2, (c) 1.70% K-2% Pt/ZrO2, (d) 2.55% K-2% Pt/ZrO2, (e) 3.40% K-2% Pt/ZrO2, (f) 4.25% K-2% Pt/ZrO2, and (g) 8.50% K-2% Pt/ZrO2.
Figure 12. TPD-MS of ethanol steam reforming over (a) 2% Pt/ZrO2, (b) 0.85% K-2% Pt/ZrO2, (c) 1.70% K-2% Pt/ZrO2, (d) 2.55% K-2% Pt/ZrO2, (e) 3.40% K-2% Pt/ZrO2, (f) 4.25% K-2% Pt/ZrO2, and (g) 8.50% K-2% Pt/ZrO2.
Nanomaterials 11 02233 g012
Figure 13. TPD-MS of ethanol steam reforming over (a) 2% Pt/ZrO2, (b) 0.37% Rb-2% Pt/ZrO2, (c) 0.74% Rb-2% Pt/ZrO2, (d) 0.93% Rb-2% Pt/ZrO2, (e) 1.86% Rb-2% Pt/ZrO2, (f) 2.79% Rb-2% Pt/ZrO2, (g) 3.72% Rb-2% Pt/ZrO2, (h) 4.65% Rb-2% Pt/ZrO2, (i) 5.58% Rb-2% Pt/ZrO2, (j) 9.29% Rb-2% Pt-ZrO2.
Figure 13. TPD-MS of ethanol steam reforming over (a) 2% Pt/ZrO2, (b) 0.37% Rb-2% Pt/ZrO2, (c) 0.74% Rb-2% Pt/ZrO2, (d) 0.93% Rb-2% Pt/ZrO2, (e) 1.86% Rb-2% Pt/ZrO2, (f) 2.79% Rb-2% Pt/ZrO2, (g) 3.72% Rb-2% Pt/ZrO2, (h) 4.65% Rb-2% Pt/ZrO2, (i) 5.58% Rb-2% Pt/ZrO2, (j) 9.29% Rb-2% Pt-ZrO2.
Nanomaterials 11 02233 g013
Figure 14. XANES and L3-L2 XANES difference spectra at the (dashed line) Pt L2 edge and (solid line) Pt L3 edge following reduction in pure hydrogen and cooling to ambient temperature, including 2% Pt/ZrO2 with: (a, blue) 0% K; (b, cyan) 0.85% K; (c, green) 1.70% K; (d, dark yellow) 2.55% K; (e, orange) 3.40% K; and (f, red) 4.25% K. (g) Overlays of L3-L2 difference spectra, showing an increase in intensity with K loading. No evidence for e- transfer to Pt from K+ was found.
Figure 14. XANES and L3-L2 XANES difference spectra at the (dashed line) Pt L2 edge and (solid line) Pt L3 edge following reduction in pure hydrogen and cooling to ambient temperature, including 2% Pt/ZrO2 with: (a, blue) 0% K; (b, cyan) 0.85% K; (c, green) 1.70% K; (d, dark yellow) 2.55% K; (e, orange) 3.40% K; and (f, red) 4.25% K. (g) Overlays of L3-L2 difference spectra, showing an increase in intensity with K loading. No evidence for e- transfer to Pt from K+ was found.
Nanomaterials 11 02233 g014
Figure 15. XANES spectra at the Pt (solid line) L3 edge and (dashed) line L2 edge, as well as (dash-dotted line) the L3-L2 difference spectra of (a, blue) 2% Pt/ZrO2, (b, cyan) 0.93% Rb-2% Pt/ZrO2, (c, green) 1.86% Rb-2% Pt/ZrO2, (d, dark yellow) 2.79% Rb-2% Pt/ZrO2, (e, orange) 4.65% Rb-2% Pt/ZrO2, (f, red) 5.58% Rb-2% Pt/ZrO2, and (g, pink) 9.3% Rb-2% Pt/ZrO2. (h) Overlays of L3 2212 L2 difference spectra, showing an increase in intensity with Rb loading. No evidence for e- transfer to Pt from Rb+ was found, which should result in an opposite trend. Figure 15 reprinted from [49] with permission from MDPI, copyright 2021.
Figure 15. XANES spectra at the Pt (solid line) L3 edge and (dashed) line L2 edge, as well as (dash-dotted line) the L3-L2 difference spectra of (a, blue) 2% Pt/ZrO2, (b, cyan) 0.93% Rb-2% Pt/ZrO2, (c, green) 1.86% Rb-2% Pt/ZrO2, (d, dark yellow) 2.79% Rb-2% Pt/ZrO2, (e, orange) 4.65% Rb-2% Pt/ZrO2, (f, red) 5.58% Rb-2% Pt/ZrO2, and (g, pink) 9.3% Rb-2% Pt/ZrO2. (h) Overlays of L3 2212 L2 difference spectra, showing an increase in intensity with Rb loading. No evidence for e- transfer to Pt from Rb+ was found, which should result in an opposite trend. Figure 15 reprinted from [49] with permission from MDPI, copyright 2021.
Nanomaterials 11 02233 g015
Figure 16. XANES derivative spectra at the Pt L3 edge of (solid line) the Pt0 foil and (red line) the catalysts following reduction in pure hydrogen, including 2% Pt/ZrO2 with: (a) 0% K; (b) 0.85% K; (c) 1.70% K; (d) 2.55% K, (e) 3.40% K; and (f) 4.25% K. No evidence for e- transfer to Pt from K+ was found.
Figure 16. XANES derivative spectra at the Pt L3 edge of (solid line) the Pt0 foil and (red line) the catalysts following reduction in pure hydrogen, including 2% Pt/ZrO2 with: (a) 0% K; (b) 0.85% K; (c) 1.70% K; (d) 2.55% K, (e) 3.40% K; and (f) 4.25% K. No evidence for e- transfer to Pt from K+ was found.
Nanomaterials 11 02233 g016
Figure 17. XANES derivative spectra at the Pt L3 edge of (solid line) the Pt0 foil and (red line) the catalysts following reduction in pure hydrogen, including 2% Pt/ZrO2 with: (a) 0% Rb, (b) 0.93% Rb, (c) 1.86% Rb, (d) 2.79% Rb, (e) 4.65% Rb, and (f) 9.3% Rb-2% Pt/ZrO2. No evidence for e- transfer to Pt from Rb+ was found.
Figure 17. XANES derivative spectra at the Pt L3 edge of (solid line) the Pt0 foil and (red line) the catalysts following reduction in pure hydrogen, including 2% Pt/ZrO2 with: (a) 0% Rb, (b) 0.93% Rb, (c) 1.86% Rb, (d) 2.79% Rb, (e) 4.65% Rb, and (f) 9.3% Rb-2% Pt/ZrO2. No evidence for e- transfer to Pt from Rb+ was found.
Nanomaterials 11 02233 g017
Table 1. Surface area, porosity, and Pt particle size for K and Rb-promoted catalysts (adapted from [48,49] with permission from Elsevier (copyright, 2020) and MDPI (copyright, 2021)).
Table 1. Surface area, porosity, and Pt particle size for K and Rb-promoted catalysts (adapted from [48,49] with permission from Elsevier (copyright, 2020) and MDPI (copyright, 2021)).
Sample IDAs (BET)
(m2/g)
Vp (BJH Des)
(cm3/g)
Dp (BJH Des)
(Å)
Est. Pt Diam
(nm)
Est. % Pt Disp.
(%)
Pt/ZrO2 (K-series)82.80.2761031.0 */0.92 **88
0.85% K-Pt/ZrO278.30.2601011.2 */1.1 **82
1.70% K-Pt/ZrO272.20.2491001.3 */1.2 **79
2.55% K-Pt/ZrO268.40.2451032.7 */2.6 **47
3.40% K-Pt/ZrO259.60.2191053.0 */3.0 **42
4.25% K-Pt/ZrO253.80.2001093.6 */3.6 **35
8.50% K-Pt/ZrO234.70.139123--
Pt/ZrO2 (Rb-series)89.70.260950.8 */0.72 **94
0.55% Rb-Pt/ZrO287.90.26896--
0.93% Rb-Pt/ZrO291.60.275930.86 */0.78 **92
1.86% Rb-Pt/ZrO288.70.262941.0 */0.93 **87
2.79% Rb-Pt/ZrO286.70.260931.1 */0.99 **85
4.65% Rb-Pt/ZrO272.30.235951.3 */1.2 **77
9.29% Rb-Pt/ZrO258.20.2021022.0 */1.9 **56
* Estimated from Jentys assuming spherical cluster morphology [50]. ** Estimated from Marinkovic et al. [51].
Table 2. Main bands in cm−1 observed for unpromoted, K-promoted, and Rb-promoted 2% Pt/ZrO2 catalysts.
Table 2. Main bands in cm−1 observed for unpromoted, K-promoted, and Rb-promoted 2% Pt/ZrO2 catalysts.
Bands0% K0.85% K2.55% K4.25% K0% Rb0.93% Rb4.65% Rb9.29% Rb
50 °C
ν(CO) ethoxy1100, 1070, 10561092, (1065), 10511103, 10581107, (1067) 10561101, 1072, 10571099, (1067), 10551098, 10561105, 1057
ν(CH) ethox/acet2970, 2928, 2897, 28682970, 2927, (2894), 28722969, 2926, (2897), 2876(2977), 2965, (2934, 2881), 28682973, 2929, (2896), 2873, (2854)2973, 2929, (2898), 2873, (2856)2973, 2931, 2898, 2879, 2858(2989), 2970, (2955), 2932, 2902, 2880, 2868
ρ(CH3) ethoxy1156, 11161148, 1124(1147, 1125)-1154, (1133), 1118(1163, 1150, 1128–1113)--
νa(OCO) acetate15621564, (1507)1577, (1519)1567–1513156415601572, (1549–1512)1578
νs(OCO) acetate(1467), 14331467, (1431), 1417(1490, 1464) 14121460(1470), 1433(1487) 1474, 1443, 14281472, (1445), 1408-
δs(CH3) acetate1381, (1357), 1344, (1274)1376, (1351) 1335, (1314–1269)1377, 1358, (1340, 1280)1358, 1310, 12711381, (1358, 1343, 1276)1381, 1357, 1343, 1327–1295, 1275(1372), 1360, (1339)1355
ν(CO) Pt-CO2055, (2037), 2012, 1990–18102051, 2030, (2018–1870)1930 (2080–1800)-2051, (2032), 2015, 1978–1870(2053, 2046–1985, 1985–1840), 1951(2063–1985, 1963), 1927, (1950–1830)-
200 °C
νa(OCO) acetate1556, (1470)(1569), 1557, (1543–1508), 14671580, (1549, 1523, 1508), 14711580, (1550–1444)(1566), 1554, (1470)1556, (1525–1470)1633, 1579, (1549, 1533–1465)1580
νs(OCO) acetate14391437, 138014081422, 140814391437, (1383)14041437, 1404
δs(CH3) acetate13461332, (1303), 1273(1351), 1334, (1297)(1350), (1330)13471339, (1330-1312, 1297, 1274)(1328), 13001355, 1339, (1335–1267)
ν(CH) acetate2965, 2937, 2876, 28623000, 2984, 2966, 2931, 28722971, 2928, (2897, 2878, 2858)(2965), 2937, 28692965, 2936, (2917–2892), 28772965, 2937, (2920–2905, 28812986, 2967, 2927(2997), 2965, (2935), 2925, (2904, 2886–2860)
500 °C
νa(OCO) carbonate15561622, 1576–14901620, (1568–1492)(1604), 1551, (1530, 1519, 1503)1550–15001580–1495, 15561627, (1581), (1550–1430)1608, (1592, 1566, 1550)
νs(OCO) carbonate(1470), 14391454(1468)(1463, 1445, 1425)1473, 14421490-1410, 1444(1550–1430)1439
νs(OCO) carbonate(1405–1367)1370, (1342) 1297, (1276–1225)(1386, 1353) 13071400-1355, 1329, (1300–1200)1395, 1363(1373), 1340, (1300), 1271, (1256)(1365, 1333, 1291)(1355), 1317
Table 3. ESR catalytic activity for K-series and Rb-series (300 °C, 1 atm, 190,560 Ncc/h/gcat, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%).
Table 3. ESR catalytic activity for K-series and Rb-series (300 °C, 1 atm, 190,560 Ncc/h/gcat, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%).
CatalystConv. C2H5OH (%)H2 Yield (%)C Selectivity (%)
CH4CO2COC2H6C2H4C3H6CH3CHO
2% Pt/ZrO286.9114.2645.2028.521.160.920.390.343.49
0.85% K-2% Pt/ZrO260.0113.8446.8640.5911.140.26--1.14
2.55% K-2% Pt/ZrO247.0212.6648.7550.72----0.76
4.55% K-2% Pt/ZrO227.303.1749.4849.08----1.42
0.93% Rb-2% Pt/ ZrO259.6013.6546.8135.7815.550.310.11-1.43
4.25% Rb-2% Pt/ZrO238.978.7053.5646.27----0.16
9.29% Rb-2% Pt/ZrO211.190.1812.1887.81-----
Table 4. ESR catalytic activity for 2% Pt/ZrO2 catalyst at different C2H5OH conversion (300 °C, 1 atm, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%).
Table 4. ESR catalytic activity for 2% Pt/ZrO2 catalyst at different C2H5OH conversion (300 °C, 1 atm, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%).
CatalystConv. C2H5OH (%)H2 Yield (%)C Selectivity (%)
CH4CO2COC2H6C2H4C3H6CH3CHO
2% Pt/ZrO286.9114.2645.2028.521.160.920.390.343.49
58.558.5347.4614.8628.370.370.570.367.98
48.414.8155.8711.5622.070.240.49-9.73
30.893.0560.797.1319.70-0.48-11.81
Table 5. ESR catalytic activity for select catalysts at the same C2H5OH conversion (300 °C, 1 atm, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%) for selectivity comparison.
Table 5. ESR catalytic activity for select catalysts at the same C2H5OH conversion (300 °C, 1 atm, feed: C2H5OH 2.98% H2O 26.14% N2 70.88%) for selectivity comparison.
CatalystConv. C2H5OH (%)H2 Yield (%)C Selectivity (%)
CH4CO2COC2H6C2H4C3H6CH3CHO
2% Pt/ZrO258.558.5347.4614.8628.370.370.570.367.98
0.85% K-2% Pt/ZrO260.0113.8446.8640.5911.140.26--1.14
0.93% Rb-2% Pt/ ZrO259.6013.6546.8135.7815.550.310.11-1.43
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martinelli, M.; Garcia, R.; Watson, C.D.; Cronauer, D.C.; Kropf, A.J.; Jacobs, G. Promoting the Selectivity of Pt/m-ZrO2 Ethanol Steam Reforming Catalysts with K and Rb Dopants. Nanomaterials 2021, 11, 2233. https://doi.org/10.3390/nano11092233

AMA Style

Martinelli M, Garcia R, Watson CD, Cronauer DC, Kropf AJ, Jacobs G. Promoting the Selectivity of Pt/m-ZrO2 Ethanol Steam Reforming Catalysts with K and Rb Dopants. Nanomaterials. 2021; 11(9):2233. https://doi.org/10.3390/nano11092233

Chicago/Turabian Style

Martinelli, Michela, Richard Garcia, Caleb D. Watson, Donald C. Cronauer, A. Jeremy Kropf, and Gary Jacobs. 2021. "Promoting the Selectivity of Pt/m-ZrO2 Ethanol Steam Reforming Catalysts with K and Rb Dopants" Nanomaterials 11, no. 9: 2233. https://doi.org/10.3390/nano11092233

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