Next Article in Journal
Nickel Supported on AlCeO3 as a Highly Selective and Stable Catalyst for Hydrogen Production via the Glycerol Steam Reforming Reaction
Next Article in Special Issue
In Silico Acetylene [2+2+2] Cycloadditions Catalyzed by Rh/Cr Indenyl Fragments
Previous Article in Journal
H3PW12O40/SBA-15 for the Solventless Synthesis of 3-Substituted Indoles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pd4S/SiO2: A Sulfur-Tolerant Palladium Catalyst for Catalytic Complete Oxidation of Methane

1
Industrial Catalysis Institute, Laboratory Breeding Base of Green Chemistry-Synthesis Technology, Zhejiang University of Technology, Hangzhou 310014, China
2
Centre for Synthesis and Chemical Biology, UCD School of Chemistry, University College Dublin, Belfield, Dublin 4, Ireland
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(5), 410; https://doi.org/10.3390/catal9050410
Submission received: 4 April 2019 / Revised: 29 April 2019 / Accepted: 29 April 2019 / Published: 30 April 2019
(This article belongs to the Special Issue Advanced Strategies for Catalyst Design)

Abstract

:
Sulfur species (e.g., H2S or SO2) are the natural enemies of most metal catalysts, especially palladium catalysts. The previously reported methods of improving sulfur-tolerance were to effectively defer the deactivation of palladium catalysts, but could not prevent PdO and carrier interaction between sulfur species. In this report, novel sulfur-tolerant SiO2 supported Pd4S catalysts (5 wt. % Pd loading) were prepared by H2S–H2 aqueous bubble method and applied to catalytic complete oxidation of methane. The catalysts were characterization by X-ray diffraction, Transmission electron microscopy, X-ray photoelectron Spectroscopy, temperature-programmed oxidation, and temperature-programmed desorption techniques under identical conditions. The structural characterization revealed that Pd4S and metallic Pd0 were found on the surface of freshly prepared catalysts. However, Pd4S remained stable while most of metallic Pd0 was converted to PdO during the oxidation reaction. When coexisting with PdO, Pd4S not only protected PdO from sulfur poisoning, but also determined the catalytic activity. Moreover, the content of Pd4S could be adjusted by changing H2S concentration of H2S–H2 mixture. When H2S concentration was 7 %, the Pd4S/SiO2 catalyst was effective in converting 96% of methane at the 400 °C and also exhibited long-term stability in the presence of 200 ppm H2S. A Pd4S/SiO2 catalyst that possesses excellent sulfur-tolerance, oxidation stability, and catalytic activity has been developed for catalytic complete oxidation of methane.

1. Introduction

Methane as the main component of natural gas is playing an increasingly important role in the global energy structure [1,2]. Effective catalytic complete oxidation of CH4 can improve combustion efficiency and reduce air pollutants, such as CO, NOx, and unburned hydrocarbon [3,4,5,6,7,8]. It has therefore found great applications in modern industry, such as catalytic exhaust converters aimed to reduce methane emission and catalytic gas turbine combustors designed to combust fuel under mild conditions [9,10]. Supported PdO catalysts have shown excellent catalytic property in methane oxidation, and currently are under extensive study [11,12,13,14,15,16,17,18]. However, once sulfur species (e.g., H2S or SO2) are present in the reaction atmosphere, the poisoning of PdO catalyst which would lead to inactive PdO-SOx is irreversible and the activity of the catalyst cannot be recovered at relatively low temperature [19,20,21,22,23,24,25].
Hence, many efforts have been devoted in the past decade to enhance sulfur resistance of supported PdO catalysts. Currently there are two primary approaches to improve the performance of palladium catalysts against sulfur poisoning: the first is to use alkaline carriers (e.g., γ-Al2O3) and enhance their adsorption ability toward acidic sulfur species to protect the active phase of PdO [23,26,27], and the second is to introduce an extra active ingredient to form a catalyst system with dual activity, such as Pd–Pt complex, to avoid sulfur poisoning [24,28]. However, these methods can only defer the deactivation of palladium catalysts as neither the alkaline carrier nor the extra active ingredient can prevent the interaction between PdO and sulfur species. In other words, sulfur poisons would eventually encroach PdO and make it inactive. As such, these catalyst systems did not really possess the ability of sulfur-tolerance.
The research on sulfur-tolerant palladium catalyst system in methane catalytic oxidation reaction is rare and lack of substantial progress in the literature [20,26,27,28]. Therefore, development of a reliable sulfur-tolerant catalyst system seems to be of great importance. In view of the great potential of catalytic oxidation of methane, we decided to pursue a new catalytic system that could be of high tolerance of sulfur poisons. We speculated PdxSy in combination with acidic carrier would be an ideal system for methane oxidation based on the following two points. First, it would be the best scenario for sulfur-tolerant catalysts that neither the support nor the palladium active phase itself could interact with sulfur species. PdxSy and acidic carrier such as SiO2 would likely meet the requirement. Second, seeing that sulfur and oxygen are located in the adjacent position of the same main group in the periodic table, i.e., the chalcogen group, the designed PdxSy might possess similar catalytic activity to PdO.
In the literature, the majority of PdxSy (Pd4S, Pd3S, Pd2.8S, Pd16S7, PdS or PdS2, etc.) were prepared by gas sulfuration with H2S–H2 [25,29,30,31,32,33,34,35]. Herein, we would like to present a new procedure for the preparation of PdxSy/SiO2 catalysts via aqueous bubble sulfuration of Pd/SiO2 with H2S–H2 with varied H2S concentration. Compared with the gas sulfuration, the aqueous bubble method was performed in much milder conditions. Besides the sulfur resistance and catalytic activity, the stability of PdxSy/SiO2 catalyst was the other focus of our investigation.

2. Results and Discussion

2.1. Fresh Catalysts

2.1.1. Catalyst Characterization

Freshly prepared catalysts were first tested by X-ray diffraction (XRD) as showed in Figure 1. The freshly prepared catalyst is named (fV), where V represents H2S concentration. No other palladium species were found except for metallic Pd0. The sulfuration of Pd to form Pd4S under H2S–H2 atmosphere was via the reaction: 4Pd + H2S Pd4S + H2 [36,37,38,39,40]. This reversible reaction meant that metallic Pd0 could not be fully sulfurized to Pd4S. Therefore, it was reasonable to detect metallic Pd0 on the surface of freshly sulfurized catalyst. However, it was worth noting that the intensity of the diffraction peak of metallic Pd0 changed regularly with the increase of H2S concentration. The average particle size of metallic Pd0 calculated by Scherrer formula was further shown in Table 1. With the increase of H2S concentration, the change of average particle size of metallic Pd0 could be divided into three stages. They were gradually reduced (0 → 5%), stable (5% → 7%), and increased (7% → 10%). These results indicated the Pd/SiO2 precursor had been changed by sulfuration.
Since XRD characterization could not directly provide evidence of the existence of PdxSy, the freshly prepared catalysts were further characterized by high resolution transmission electron microscopy (HR-TEM) and illustrated in Figure 2. The lattice spacing distance of Pd species particles on freshly prepared catalyst are 0.225 nm and 0.245 nm, corresponding to the Pd(111) and Pd4S(102), respectively [JCPDS No. 73-1387, JCPDS No. 46-1043]. In addition to the metallic Pd(111), the Pd4S(102) was found on the surface of all the freshly prepared catalysts. The literature on the preparation of PdxSy under H2S-H2 atmosphere showed that the Pd4S with the highest proportion of Pd/S was the first PdxSy species [38,41]. Therefore, it was reasonable to firstly form the Pd4S under the condition of 60 °C aqueous bubble sulfuration.
The subsequent characterization by X-ray photoelectron spectroscopy (XPS) could provide further evidence for the existence of PdxSy, as shown in Figure 3. After curve fitting analysis, the S2p3/2 of freshly prepared catalyst could be deconvoluted into a large peak at 168.5 eV and a small peak at 165.1 eV, which corresponded to S0 [42] and Pd4S [42,43,44], respectively. With the increase of H2S concentration, the XPS peak intensity of sulfur species increased gradually. At the same time, the Pd3d5/2 could be deconvoluted into two peaks at ~335.6 eV and ~337.5 eV which corresponded to metallic Pd0 [45,46] and Pd4S [47,48], respectively. This indicated that Pd4S and metallic Pd0 were the primary palladium species on the freshly prepared catalysts. These results are in good agreement with the characterization results of XRD and HR-TEM. If the Pd4S had similar properties to PdO, the amount of Pd4S would be the key factor influencing the performance of the catalyst. The XPS data (Figure 3) were further analyzed to provide Table 2 that listed the composition ratio of palladium species of freshly prepared catalysts. It could be noted that the sulfuration process was the transition of metallic Pd0 to Pd4S. With the increase of H2S concentration, the change of Pd4S ratio could also be divided into three stages. They were gradually increased (0 → 5%), stable (5% → 7%), and reduced (7% → 10%). Variation of metallic Pd0 content was opposite to that of Pd4S. This variation was entirely consistent with the average particle size of metallic Pd0. It was obvious that the variation of the average particle size of metallic Pd0 was caused by the change of metallic Pd0 content on the catalyst surface.

2.1.2. Activity Studies

In order to test whether the Pd4S possessed the desired sulfur-tolerance and catalytic activity, its catalytic property for methane oxidation in the presence of 200 ppm H2S was then examined by measuring methane conversion against the reaction time at different reaction temperatures. As shown in Figure 4, when the temperature was 500 °C, the presence of H2S had no impact on all the catalysts. However, once the reaction temperature dropped to 400 °C, the compared PdO/SiO2 catalyst was deactivated very rapidly. We estimated that at this temperature sulfur species could turn the active PdO into inactive PdO–SOx species, which could then hardly decompose at low temperature [23,24,49]. In contrast, under the same conditions, all the catalysts prepared by our own procedure did not show any deactivation at 400 °C. This clearly stated that the combination of Pd4S and SiO2 was effective in resisting sulfur poisoning. At the same time, the activity of the catalyst with 7% H2S concentration was even equal to that of PdO catalyst. This further stated that the combination of Pd4S and SiO2 had high enough catalytic activity. Furthermore, it was found that the methane conversion of all catalysts decreased to less than 10% at 370 °C. This dramatic decrease in catalytic activity was not due to sulfur poisoning, but the high reaction temperature required for methane activation [50,51,52]. Therefore, it was best to set the reaction temperature at 400 °C. This temperature could not only reflect the sulfur-tolerance, but also compare the catalytic activity.
Meanwhile, it should be pointed out here that although all the sulfurized catalysts exhibited the similar sulfur-tolerance, the catalytic activity at 400 °C was not same. As the H2S concentration increased from 0.2% to 7%, the catalytic activity increased gradually. However, when the H2S concentration increased to 10%, the catalytic activity decreased. This variation was very similar to the variation of Pd4S content on the surface of freshly prepared catalyst. However, we could not simply assume that catalytic activity depended solely on Pd4S. The reason was that the stability of Pd4S and metallic Pd0 on the surface of freshly prepared catalyst was still undiscovered under high reaction temperature and presence of oxygen. Under such reaction condition, metallic Pd0 was liable to be oxidized to PdO. Therefore, palladium species on the surface of freshly prepared catalyst might not be the true palladium species involved in the reaction. Next, we would characterize the used catalyst to determine the actual palladium species on the catalyst surface.

2.2. Used Catalyst

2.2.1. Catalyst Characterization

The XRD patterns of used catalysts were given in Figure 5. The used catalyst is named (uV), where V represents H2S concentration. The characteristic diffraction peak attributed to sulfur species was found on the compared PdO/SiO2 catalyst. However, no diffraction peaks attributed to sulfur species were found in all sulfurized catalysts after the reaction. The reason should be that the sulfurized catalyst had the ability to resist sulfur poisoning. Not surprisingly, almost all the diffraction peaks observed in the sulfurized catalyst were attributed to PdO after the reaction. A weak diffraction peak of metallic Pd0 could be observed only when H2S concentration was 0.2%. This implied that most of the metallic Pd0 on the freshly prepared catalyst had been oxidized to PdO under reaction conditions. However, the above results did not show whether the most important Pd4S remains stable during the reaction and this needed to be identified by other characterization.
The presence of PdO confirmed our previous speculation that the palladium species on the surface of freshly prepared catalysts were not entirely stable during the vigorous oxidation reaction. However, XRD characterization could only reveal that PdO species were formed during the reaction of freshly prepared catalysts. However, whether the essential Pd4S species could stably exist in large quantities would be the main factor affecting the sulfur-tolerance of the catalyst. Figure 6 shows HR-TEM images of the used catalysts. It could be found that the palladium species on the surface of the used catalyst were more complicated than the freshly prepared catalyst. Figure 6 clearly indicates that in addition to Pd0(111), PdO(101) and PdO(110) also appeared in large numbers on the surface of all the used catalysts [JCPDS No.06-0515]. At the same time, the most interesting thing was that Pd4S(102) species could be found on all catalyst surfaces. This indicated that the Pd4S species could remain stable during the high temperature and high oxygen reaction of methane oxidation.
Owing to the lack of data on the relative content of PdO on the surface of used catalyst, when Pd4S and PdO coexist, we could not distinguish whether Pd4S and PdO affect the catalytic activity alone or in combination. Therefore, used catalysts were further analyzed by XPS characterization, as shown in Figure 7. After curve fitting analysis, the Pd3d5/2 could be deconvoluted into three peaks at ~337.5 eV, ~336.9 eV, and ~335.6 eV, which were assigned to Pd4S [47,48], PdO [48], and metallic Pd0 [45,46], respectively. The above characterization results revealed the fact that under the reaction conditions, the actual palladium species on the used catalyst surface should be Pd4S, PdO, and metal Pd0. At the same time, the S2p peak attributed to S0 was not observed (see Figure 3), only the very weak S2p peak attributed to Pd4S could be observed. In contrast, the Pd3d peak at ~337.3 eV and S2p peak at 168.1 eV of used PdO/SiO2 could be attributed to PdO–SOx. This comparison results indicated that Pd4S could protect other palladium species from sulfur poisoning.
Table 3 listed the composition ratio of palladium species of used catalysts. Compared with the relative content of palladium species on the surface of fresh catalyst (Table 1), it could be noted that a large number of PdO species appeared on the surface of the used catalyst, while the Pd0 content of the metal decreased significantly. This was clearly the oxidation of metal Pd0 to PdO during the reaction. At the same time, the relative content of Pd4S was basically stable during the reaction. This indicated that the PdO formed during the reaction was mainly derived from the metal Pd0, but not derived from Pd4S. Temperature-programmed Oxidation (TPO) of fresh Pd/SiO2, f7 and u7 catalysts with the same mass was investigated by thermogravimetric analysis. The oxidation stability of the catalyst was examined by recording the mass change in the process of air oxidation. The results are presented in Figure 8. Three samples had a similar dehydration process before 100 °C. However, with the increase of temperature, the three samples showed completely different changes. Freshly prepared Pd/SiO2 had the greatest mass increase, which should be related to the oxidation of metallic Pd0 to PdO. The mass increase of f7 with the same mass was significantly smaller than that of freshly prepared Pd/SiO2, while u7 with the same mass had no significant mass increase. f7 was composed of metallic Pd0 and Pd4S, while u7 was mainly composed of PdO and Pd4S. This stated that the mass increase of f7 was mainly related to the oxidation of metallic Pd0, while Pd4S maintained sufficient oxidation stability even under temperature higher than 500 °C. This meant that Pd4S could not be oxidized into palladium oxide nor into palladium sulfate. Therefore, it could be possible to conclude that in the low-temperature catalytic oxidation of methane, Pd4S had sufficient stability to resist the oxidation of oxygen.

2.2.2. Activity Studies

Figure 9 covered the evaluation of the used catalysts and the evaluation method was completely consistent with the freshly prepared catalysts. To our delight, the sulfur-tolerance and catalytic activity of each used catalyst was almost unchanged with the fresh catalyst at 400 °C, which indicated that the composition of the catalyst had remained stable after the initial evaluation. The previous characterization results had confirmed that Pd4S, PdO, and metal Pd0 were the three palladium species actually involved in the reaction. In the case that PdO and metal Pd0 did not have sulfur-tolerance, it was the existence of Pd4S that gave the palladium active component the ability to resist the poisoning of sulfur species. Moreover, to our surprise, Pd4S was not merely resistant to sulfur itself, its presence could even protect PdO and metal Pd0 from the poisoning of sulfur species.
The sulfur-tolerance of the catalyst depended on Pd4S, but it remained to be confirmed if catalytic activity was related to the type of palladium species. To this end, we correlated the relative amounts of Pd4S and PdO species on the used catalyst surface with the catalytic activity of the used catalyst. The secondary reaction performance of the catalyst was selected at 400 °C and the results are shown in Figure 10. It could be noted that the variation of catalytic activity at 400 °C was closely related to the change of Pd4S content with H2S concentration increase. As the H2S concentration increased from 0.2% to 7%, the relative content of Pd4S and the catalytic activity of the used catalyst gradually increased. Compared with content change of (Pd4S + PdO) and PdO, the content of (Pd4S + PdO) did not change with the concentration of H2S, but the change of PdO content was opposite to the catalytic activity of the used catalyst. The above results showed clearly that not only the sulfur-tolerance, but also the catalytic activity relied completely on Pd4S. Because the coexistence of PdO did not play any decisive role, it was completely reasonable to use Pd4S/SiO2 as the catalyst prepared by the aqueous bubble sulfuration method. The consequences of 72-hour stability test of the u7 catalyst with and without H2S are shown Figure 11. It could be found that the catalyst could maintain long-term stability regardless of the presence of H2S in the reaction atmosphere.

2.3. Mechanism of Sulfur-Tolerance

Next, we turned to investigate the cause of sulfur-tolerance of the Pd4S/SiO2 using the technique of hydrogen sulfide temperature-programmed desorption (H2S-TPD). The u7 catalyst and the freshly prepared PdO/SiO2 catalyst were chosen for comparison and the results are presented in Figure 12. SO2 and H2S were the primary desorption species detected in the study. The PdO/SiO2 catalyst released SO2 in relatively large quantity, which suggested that PdO could readily assimilate H2S to form PdO–SOx. As a result, PdO/SiO2 catalyst was poisoned and deactivated by H2S. In contrast, desorption products were barely detected from the u7 catalyst. Seeing that Pd4S was the only discrepancy between the two catalysts, i.e., the presence of Pd4S in u7 and its absence in PdO/SiO2, we were convinced that Pd4S played a vital role in the sulfur-tolerance of u7 catalyst, as depicted in Scheme 1. Aside from not adsorbing H2S, Pd4S also prevented PdO from adsorbing H2S onto its surface. Therefore, it was the Pd4S that blocked the adsorption of H2S and endowed the catalyst with sulfur-tolerance.

3. Materials and Methods

3.1. Catalyst Preparation

The PdO/SiO2 catalyst was prepared by isovolumetric impregnation. Typically, weigh a proper quantity of 40~60 mesh SiO2 (M = 10 g Vg = 0.94 cm3/g, Rd = 10.6 nm, S = 353.0028 m2/g, Qingdao Baisha Catalyst Factory, Qingdao, China) was dispersed in palladium acetate (Aladdin Industrial Corporation, Shanghai, China) aqueous solution (VPd = 0.05 g/mL) overnight. Then the sample was dried in an oven at 110 °C for 4 h and calcined at 500 °C for 4 h in the air. The theoretical loading of Pd was 5 wt. %.
The PdO/SiO2 was first reduced to Pd/SiO2 by H2 at 500 °C for 1h. PdxSy/SiO2 catalyst was prepared by using above Pd/SiO2 as precursor. 2.0 g Pd/SiO2 and 100 ml deionized water were stirred at 350 rpm in a three-necked flask. Adjust the flow rate of H2S and H2 by flow controller (D07-19B, Beijing Sevenstart Electronics Co. Ltd, Beijing, China) to configure different concentrations of H2S/H2 mixture. Then 30 mL/min of H2S–H2 (VH2S% = 0.2, 1, 3, 5, 7, 10) was fed into the suspension at 60 °C for 1 h. Afterwards, the sample was filtered and rinsed with distilled water till neutral. Finally, the sample was dried at 110 °C for 4 h. The freshly prepared catalyst is named (fV), and the used catalyst is named (uV), where V represents H2S concentration.

3.2. Catalyst Characterization

X-Ray Diffraction (XRD) was carried out on a Thermo ARL X’TRA diffractometer (PNAlytical Co. Holland) using Cu K-a radiation (45 kV, 40 mA). Average particle size was determined from XRD measurements using the Scherrer formula:
d = K λ β cos θ ,
where d is the average particle size(nm), K is the Scherrer constant and the diffraction angle is denoted θ, λ = 0.154 nm stands for the wavelength of Cu K-a radiation, β denotes the full-width at half-maximum (FWHM) of diffraction peak.
High Resolution Transmission Electron Microscopy (HR-TEM) images were taken by a Philips-FEI Tecnai G2 F30 S-Twin transmission electron microscope operated at 300 kV (Philips-FEI Co. Holland).
X-Ray Photoelectron Spectroscopy (XPS) was measured on a Thermo ESCALAB 250 Axis Ultra (KRATOS, Kanagawa, Japan) using a monochromated Al Kα X-ray source (hv = 1485.6 eV) with a fixed analyzer pass energy of 80 eV. The binding energy values were referenced to the Si 2p as internal standard (Si 2p = 103.4 eV) and the maximum deviation value was 2.6 eV in the sample. After subtraction of the Shirley-type background, the core-level spectra were decomposed into their component with mixed Gaussian-Lorentzan lines by a non-linear least squares curve fitting procedure, using the public software package XPSPEAK4.1. The corresponding atomic ratio in different chemical environment was obtained from the fitted XPS spectra of Pd 3d and S 2p. The XPS profiles were fitted by the software named “XPS peak”. Then the fitting peak area was used to calculate the proportion of different species.
Temperature-programmed Oxidation (TPO) was performed on a NETZSCH STA 409 PC/PG instrument (NETZSCH Co. Selbc, Germany) The oxidation stability of the catalysts was investigated by recording the mass change in the process of air oxidation. TPO was performed by heating 0.01 g sample from room temperature to 600 °C in air atmosphere with a heating rate of 5 °C/min and a flow rate of 40 mL/min.
Temperature-programmed Desorption (TPD) of hydrogen sulfide experiments were performed in a self-made tubular quartz reactor (5mm i.d.). The sample (0.2 g) was first swept at 110 °C for 1 h using pure He and cooled to room temperature in the same atmosphere. Then the sample was in situ treated with H2S (0.2% in N2) at a flow rate of 30 mL/min for 0.5 h and swept 1 h to remove physisorbed and/or weakly bound species. TPD was performed by heating the sample from room temperature to 800 °C with constant increase of 5 °C/min in pure He. The TPD spectra were recorded by a quadrupole mass spectrometer (QMS 200 Omnistar, Pfeiffer Co. Germany).

3.3. Evaluation of Catalysts

The catalytic activity of the freshly prepared and used catalysts was tested with 0.2 g sample in a self-made continuous fixed bed reactor (8 mm i.d.) at atmospheric pressure. Gases consisting of CH4 (v% = 2 %), O2 (v% = 8 %), H2S (v% = 0.02 %), and N2 were fed through the flow controller (D07-19B, Beijing Sevenstart Electronics Co. Ltd., Beijing, China) at 200 mL/min and used in all the experiments. Finally, the temperature control controller (AI808PK1L2, XiaMen YuDian Automation Technology Co., Ltd., XiaMen, China) is used to control different reaction temperatures. The effluent gases were sampled and simultaneously analyzed online by gas chromatographs (GC-9790, Zhejiang Fuli Analytical Instruments Corp., Hangzhou, China). Exhaust gases were analyzed using a PoraPak Q column and a thermal conductivity detector (TCD). Methane conversions were calculated for outlet CO2 concentrations.

4. Conclusions

Through the characterization of freshly prepared catalyst and used catalyst, we confirmed that Pd4S, PdO, and metallic Pd0 were the actual palladium species involved in the oxidation of methane. Sulfur-tolerance and catalytic activity was completely dependent on Pd4S, and independent on other palladium species. Moreover, Pd4S had sufficient oxidation stability under reaction conditions. Therefore, even in the presence of PdO, there are enough reasons to represent the catalyst with Pd4S/SiO2. In summary, we have developed a powerful sulfur-tolerant catalyst for methane oxidation by incorporating Pd4S into SiO2, and to the best of our knowledge, this is the first report on Pd catalysts with such property. In view of the excellent sulfur-tolerant property, the excellent oxidation stability and the high catalytic activity, the Pd4S/SiO2 catalyst will definitely find valuable and versatile use in catalytic chemistry.
At the same time, as the PdxSy species with the highest proportion of Pd/S, we cannot determine whether the outstanding performance of Pd4S in methane catalytic oxidation is only a special case. Furthermore, sulfur-tolerant mechanism of Pd4S needs further theoretical research.

Author Contributions

Conceptualization, L.M.; Data curation, S.Y. and H.Z.; Formal analysis, L.M., S.Y., and H.Z.; Funding acquisition, L.M. and C.L.; Investigation, T.J.; Methodology, X.Z.; Project administration, L.M. and X.L.; Writing—original draft, S.Y.; Writing—review & editing, L.M. and X.Z.

Funding

This work is financially supported by National Natural Science Foundation of China (Grant No.21473159, 21476208).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trinchero, A.; Hellman, A.; Grönbeck, H. Methane oxidation over Pd and Pt studied by DFT and kinetic modeling. Surf. Sci. 2013, 616, 206–213. [Google Scholar] [CrossRef]
  2. Mahara, Y.; Ohyama, J.; Tojo, T.; Murata, K.; Ishikawa, H.; Satsuma, K. Enhanced activity for methane combustion over a Pd/Co/Al2O3 catalyst prepared by a galvanic deposition method. Catal. Sci. Technol. 2016, 6, 4773–4776. [Google Scholar] [CrossRef]
  3. Chen, J.; Arandiyan, H.; Gao, X.; Li, J. Recent Advances in Catalysts for Methane Combustion. Catal. Surv. Asia 2015, 19, 140–171. [Google Scholar] [CrossRef]
  4. Colussi, S.; Gayen, A.; Camellone, M.F.; Boaro, M.; Llorca, J.; Fabris, S.; Trovarelli, A. Nanofaceted Pd-O Sites in Pd-Ce Surface Superstructures: Enhanced Activity in Catalytic Combustion of Methane. Angew. Chem. 2009, 121, 8633–8636. [Google Scholar] [CrossRef]
  5. Ji, Y.; Guo, Y.B. Nanostructured perovskite oxides as promising substitutes of noble metals catalysts for catalytic combustion of methane. Chin. Chem. Lett. 2018, 29, 252–260. [Google Scholar]
  6. Bossche, M.V.D.; Gronbeck, H. Methane Oxidation over PdO(101) Revealed by First-Principles Kinetic Modeling. J. Am. Chem. Soc. 2015, 137, 12035–12044. [Google Scholar] [CrossRef] [PubMed]
  7. Monai, M.; Montini, T.; Gorte, R.J.; Fornasiero, P. Catalytic Oxidation of Methane: Pd and Beyond. Eur. J. Inorg. Chem. 2018, 25, 2884–2893. [Google Scholar] [CrossRef]
  8. Jiang, L.; Zheng, Y.; Chen, X.; Xiao, Y.; Cai, G.; Zheng, Y.; Zhang, Y.; Huang, F. Catalytic Activity and Stability over Nanorod-Like Ordered Mesoporous Phosphorus-Doped Alumina Supported Palladium Catalysts for Methane Combustion. ACS Catal. 2018, 8, 11016–11028. [Google Scholar]
  9. Fouladvand, S.; Skoglundh, M.; Carlsson, P.A. A transient in situ infrared spectroscopy study on methane oxidation over supported Pt catalysts. Catal. Sci. Technol. 2014, 4, 3463–3473. [Google Scholar] [CrossRef]
  10. Florén, C.R.; Bossche, M.V.D.; Creaser, D.; Grobeck, H.; Carlsson, P.A.; Korpi, H.; Skoglundh, M. Modelling complete methane oxidation over palladium oxide in a porous catalyst using first-principles surface kinetics. Catal. Sci. Technol. 2018, 8, 508–520. [Google Scholar] [CrossRef]
  11. Banerjee, A.C.; Golub, K.W.; Abdul, Md. H.; Billor, M.Z. Comparative study of the characteristics and activities of Pd/γ-Al2O3 catalysts prepared by Vortex and Incipient Wetness Methods. Catalysts 2019, 9, 336. [Google Scholar] [CrossRef]
  12. Banerjee, A.C.; McGuire, M.M.; Lawnick, O.; Bozack, M.J. Low-temperature activity and PdO/PdOx transition in methane combustion by a PdO-PdOx/γ-Al2O3 catalyst. Catalysts 2018, 8, 266. [Google Scholar] [CrossRef]
  13. Ciuparu, D.; Lyubovsky, M.R.; Altman, E.; Pfefferle, L.D.; Datye, A. Catalytic combustion of methane over palladium based catalysts. Catal. Rev. Sci. Eng. 2002, 44, 593–649. [Google Scholar] [CrossRef]
  14. Persson, K.; Pfefferle, L.D.; Schwartz, W.; Ersson, A.; Jaras, S.G. Stability of palladium-based catalysts during catalytic combustion of methane: The influence of water. Appl. Catal. B 2007, 74, 242–250. [Google Scholar] [CrossRef]
  15. Liu, Y.; Wang, S.; Gao, D.; Sun, T.; Zhang, C.; Wang, S.D. Influence of metal oxides on the performance of Pd/Al2O3 catalysts for methane combustion under lean-fuel conditions. Fuel Process. Technol. 2013, 111, 55–61. [Google Scholar] [CrossRef]
  16. Baylet, A.; Royer, S.; Marecot, P.; Tatibouet, J.M.; Duprez, D. High catalytic activity and stability of Pd doped hexaaluminate catalysts for the CH4 catalytic combustion. Appl. Catal. B 2008, 77, 237–247. [Google Scholar] [CrossRef]
  17. Ma, J.; Lou, Y.; Cai, Y.; Zhao, Z.; Wang, L.; Zhan, W.; Guo, Y.; Guo, Y. The relationship between the chemical state of Pd species and the catalytic activity for methane combustion on Pd/CeO2. Catal. Sci. Technol. 2018, 8, 2567–2577. [Google Scholar] [CrossRef]
  18. Mihai, O.; Smedler, G.; Nylén, U.; Olofsson, M.; Olsson, L. The effect of water on methane oxidation over Pd/Al2O3 under lean, stoichiometric and rich conditions. Catal. Sci. Technol. 2017, 7, 3084–3096. [Google Scholar] [CrossRef]
  19. Zi, X.; Liu, L.; Xue, B.; Dai, H.; He, H. The durability of alumina supported Pd catalysts for the combustion of methane in the presence of SO2. Catal. Today 2011, 175, 223–230. [Google Scholar] [CrossRef]
  20. Wilburn, M.S.; Epling, W.S. Sulfur deactivation and regeneration of mono- and bimetallic Pd-Pt methane oxidation catalysts. Appl. Catal. B 2017, 206, 589–598. [Google Scholar] [CrossRef]
  21. Monai, M.; Montini, T.; Melchionna, M.; Ducchon, T.; Kus, P.; Chen, C.; Tsud, N.; Nasi, L.; Prince, K.C.; Veltruska, K.; et al. The effect of sulfur dioxide on the activity of hierarchical Pd-based catalysts in methane combustion. Appl. Catal. B 2017, 202, 72–83. [Google Scholar] [CrossRef]
  22. Yin, F.; Ji, S.; Wu, P.; Zhao, F.; Liu, H.; Li, C. Preparation of Pd-Based Metal Monolithic Catalysts and a Study of Their Performance in the Catalytic Combustion of Methane. ChemSusChem 2008, 1, 311–319. [Google Scholar] [CrossRef]
  23. Hoyos, L.J.; Praliaud, H.; Primet, M. Catalytic combustion of methane over palldium supported on alumina and silica in presence of hydrogen-sulfie. Appl. Catal. A 1993, 98, 125–138. [Google Scholar] [CrossRef]
  24. Venezia, A.M.; Carlo, G.D.; Liotta, L.F.; Pantaleo, G.; Kantcheva, M. Effect of Ti(IV) loading on CH4 oxidation activity and SO2 tolerance of Pd catalysts supported on silica SBA-15 and HMS. Appl. Catal. B 2011, 106, 529–539. [Google Scholar] [CrossRef]
  25. Ortloff, F.; Bohnau, J.; Kramar, U.; Graf, F.; Kolb, T. Studies on the influence of H2S and SO2 on the activity of a PdO/Al2O3 catalyst for removal of oxygen by total oxidation of (bio-)methane at very low O2:CH4 ratios. Appl. Catal. B 2016, 182, 550–561. [Google Scholar] [CrossRef]
  26. Meeyoo, V.; Trimm, D.L.; Cant, N.W. The effect of sulfur containing pollutants on the oxidation activity of precious metals used in vehicle exhaust catalysts. Appl. Catal. B 1998, 16, L101–L104. [Google Scholar] [CrossRef]
  27. Ordóñez, S.; Hurtado, P.; Díez, F.V. Methane catalytic combustion over Pd/Al2O3 in presence of sulfur dioxide: development of a regeneration procedure. Catal. Lett. 2005, 100, 27–34. [Google Scholar] [CrossRef]
  28. Corro, G.; Cano, C.; Fierro, J.L.G. A study of Pt–Pd/γ-Al2O3 catalysts for methane oxidation resistant to deactivation by sulfur poisoning. J. Mol. Catal. A: Chem 2010, 315, 35–42. [Google Scholar] [CrossRef]
  29. Simon, L.J.; Ommen, J.G.V.; Jentys, A.; Lercher, J.A. Sulfur-Tolerant Pt-Supported Zeolite Catalysts for Benzene Hydrogenation: I. Influence of the Support. J. Catal. 2001, 201, 60–69. [Google Scholar] [CrossRef]
  30. Ferrer, I.J.; Diazchao, P.; Pascual, A.; Sánchez, C. An investigation on palladium sulphide (PdS) thin films as a photovoltaic material. Thin Solid Films 2007, 515, 5783–5786. [Google Scholar] [CrossRef]
  31. Xu, W.; Ni, J.; Zhang, Q.F.; Feng, F.; Xiang, Y.Z.; Li, X.N. Tailoring supported palladium sulfide catalysts through H2-assisted sulfidation with H2S. J. Mater. Chem. A 2013, 1, 12811–12817. [Google Scholar] [CrossRef]
  32. Zhang, Q.F.; Xu, W.; Li, X.N.; Jiang, D.H.; Xiang, Y.Z.; Wang, J.G.; Cen, J.; Romano, S.; Ni, J. Catalytic hydrogenation of sulfur-containing nitrobenzene over Pd/C catalysts: In situ sulfidation of Pd/C for the preparation of PdxSy catalysts. Appl. Catal. A 2015, 497, 17–21. [Google Scholar] [CrossRef]
  33. Mccue, A.J.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I.; Anderson, J.A. Palladium sulphide – A highly selective catalyst for the gas phase hydrogenation of alkynes to alkenes. J. Catal. 2016, 340, 10–16. [Google Scholar] [CrossRef]
  34. Menegazzo, F.; Canton, P.; Pinna, F.; Pernicone, N. Bimetallic Pd–Au catalysts for benzaldehyde hydrogenation: Effects of preparation and of sulfur poisoning. Catal. Commun. 2008, 9, 2353–2356. [Google Scholar] [CrossRef]
  35. Mccue, A.J.; Anderson, J.A. Sulfur as a catalyst promoter or selectivity modifier in heterogeneous catalysis. Cheminform 2014, 4, 272–294. [Google Scholar] [CrossRef]
  36. O’Brien, C.P.; Gellman, A.J.; Morreale, B.D.; Miller, J.B. The hydrogen permeability of Pd4S. J. Membr. Sci. 2011, 371, 263–267. [Google Scholar] [CrossRef]
  37. O’Brien, C.P.; Howard, B.H.; Miller, J.B.; Morreale, B.D.; Gellman, A.J. Inhibition of hydrogen transport through Pd and Pd47Cu53 membranes by H2S at 350 degrees. J. Membr. Sci. 2010, 349, 380–384. [Google Scholar] [CrossRef]
  38. Morreale, B.D.; Howard, B.H.; Iyoha, O.; Enick, R.M.; Ling, C.; Sholl, D.S. Experimental and Computational Prediction of the Hydrogen Transport Properties of Pd4S. Ind. eng. chem. res 2007, 46, 6313–6319. [Google Scholar] [CrossRef]
  39. Zubkov, A.; Fujino, T.; Sato, N.; Yamada, K. Enthalpies of formation of the palladium sulphides. J. Chem. Thermodyn. 1998, 30, 571–581. [Google Scholar] [CrossRef]
  40. Iyoha, O.; Enick, R.; Killmeyer, R.; Morreale, B. The influence of hydrogen sulfide-to-hydrogen partial pressure ratio on the sulfidization of Pd and 70 mol% Pd–Cu membranes. J. Membr. Sci. 2007, 305, 77–92. [Google Scholar] [CrossRef]
  41. Hensen, E.J.M.; Brans, H.J.A.; Lardinois, G.M.H.J.; deBeer, V.H.J.; vanVeen, J.A.R.; van Santen, R.A. Periodic Trends in Hydrotreating Catalysis: Thiophene Hydrodesulfurization over Carbon-Supported 4d Transition Metal Sulfides. J. Catal. 2000, 192, 98–107. [Google Scholar] [CrossRef]
  42. Gerson, A.R.; Bredow, T. Interpretation of sulfur 2p XPS spectra in sulfide minerals by means of ab initio calculations. Surf. Interface Anal. 2000, 29, 145–150. [Google Scholar] [CrossRef]
  43. Bhatt, R.; Bhattacharya, S.; Basu, R.; Singh, A.; Deshpande, U.; Surger, C.; Basu, S.; Aswal, D.K.; Gupta, S.K. Growth of Pd4S, PdS and PdS2 films by controlled sulfurization of sputtered Pd on native oxide of Si. Thin Solid Films 2013, 539, 41–46. [Google Scholar] [CrossRef]
  44. Senftle, T.P.; Van Duin, A.C.T.; Janik, M.J. The Role of Site Stability in Methane Activation on PdxCe1-xOδ Surfaces. ACS Catal. 2015, 5, 6187–6199. [Google Scholar] [CrossRef]
  45. Beketov, G.; Heinrichs, B.; Pirard, J.P.; Chenakin, S.; Kruse, N. XPS structural characterization of Pd/SiO2 catalysts prepared by cogelation. Appl. Surf. Sci. 2013, 287, 293–298. [Google Scholar] [CrossRef]
  46. Romanchenko, A.S.; Mikhlin, Y.L. An XPS study of products formed on pyrite and pyrrhotine by reacting with palladium (II) chloride solutions. J. Struct. Chem. 2015, 56, 531–537. [Google Scholar] [CrossRef]
  47. Corro, G.; Vázquez-Cuchillo, O.; Banuelos, F.; Fierro, J.L.G.; Azomoza, M. An XPS evidence of the effect of the electronic state of Pd on CH4 oxidation on Pd/gamma-Al2O3 catalysts. Catal. Commun. 2007, 8, 1977–1980. [Google Scholar] [CrossRef]
  48. Chenakin, S.P.; Melaet, G.; Szukiewicz, R.; Kruse, N. XPS study of the surface chemical state of a Pd/(SiO2+TiO2) catalyst after methane oxidation and SO2 treatment. J. Catal. 2014, 312, 1–11. [Google Scholar] [CrossRef]
  49. Weng, X.; Ren, H.; Chen, M.; Wan, H. Effect of Surface Oxygen on the Activation of Methane on Palladium and Platinum Surfaces. ACS Catal. 2014, 4, 2598–2604. [Google Scholar] [CrossRef]
  50. Cargnello, M.; Jaén, J.J.D.; Garrido, J.C.H.; Bakhmutsky, K.; Montini, T.; Gámez, J.J.C.; Gorte, R.J.; Fornasiero, P. Exceptional Activity for Methane Combustion over Modular Pd@CeO2 Subunits on Functionalized Al2O3. Science 2012, 337, 713–718. [Google Scholar] [CrossRef] [PubMed]
  51. Widjaja, H.; Sekizawa, K.; Eguchi, K.; Arai, H. Oxidation of methane over Pd/mixed oxides for catalytic combustion. Catal. Today 1999, 47, 95–101. [Google Scholar] [CrossRef]
  52. Sekizawa, K.; Widjaja, H.; Maeda, S.; Ozawa, Y.; Eguchi, K. Low temperature oxidation of methane over Pd catalyst supported on metal oxides. Catal. Today 2000, 59, 69–74. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of freshly prepared catalysts sulfurized with different H2S concentration.
Figure 1. X-ray diffraction (XRD) patterns of freshly prepared catalysts sulfurized with different H2S concentration.
Catalysts 09 00410 g001
Figure 2. High resolution transmission electron microscopy (HR-TEM) images of freshly prepared catalysts with different H2S concentration.
Figure 2. High resolution transmission electron microscopy (HR-TEM) images of freshly prepared catalysts with different H2S concentration.
Catalysts 09 00410 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) Pd3d and S2p of freshly prepared catalysts with different H2S concentration.
Figure 3. X-ray photoelectron spectroscopy (XPS) Pd3d and S2p of freshly prepared catalysts with different H2S concentration.
Catalysts 09 00410 g003
Figure 4. Catalytic performance of freshly prepared catalysts with different H2S concentration. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 89.8%); Flow rate = 200 mL/min; and GHSV (gas hourly space velocity) = 60000 mL/(g·h).
Figure 4. Catalytic performance of freshly prepared catalysts with different H2S concentration. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 89.8%); Flow rate = 200 mL/min; and GHSV (gas hourly space velocity) = 60000 mL/(g·h).
Catalysts 09 00410 g004
Figure 5. XRD patterns of used catalysts with different H2S concentration.
Figure 5. XRD patterns of used catalysts with different H2S concentration.
Catalysts 09 00410 g005
Figure 6. HR-TEM images of the used catalysts with different H2S concentration.
Figure 6. HR-TEM images of the used catalysts with different H2S concentration.
Catalysts 09 00410 g006
Figure 7. XPS Pd3d and S2p of the used catalysts with different H2S concentration.
Figure 7. XPS Pd3d and S2p of the used catalysts with different H2S concentration.
Catalysts 09 00410 g007
Figure 8. Temperature-programmed Oxidation (TPO) profiles of the (u7, f7) catalyst and freshly prepared Pd/SiO2.
Figure 8. Temperature-programmed Oxidation (TPO) profiles of the (u7, f7) catalyst and freshly prepared Pd/SiO2.
Catalysts 09 00410 g008
Figure 9. Catalytic performance of used catalysts with different H2S concentration. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 8 9.8%); Flow rate: 200 mL/min; and GHSV = 60000 mL/(g·h).
Figure 9. Catalytic performance of used catalysts with different H2S concentration. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 8 9.8%); Flow rate: 200 mL/min; and GHSV = 60000 mL/(g·h).
Catalysts 09 00410 g009
Figure 10. Relationship profile between composition ratio of Pd4S and PdO on used catalysts and second evaluation of methane conversion. The composition ratio of palladium species was calculated through Pd fitted peak area by XPS.
Figure 10. Relationship profile between composition ratio of Pd4S and PdO on used catalysts and second evaluation of methane conversion. The composition ratio of palladium species was calculated through Pd fitted peak area by XPS.
Catalysts 09 00410 g010
Figure 11. Long-term stability test of the u7 catalyst with and without H2S. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 89.8%); Flow rate: 200 mL/min; and GHSV = 60000 mL/(g·h).
Figure 11. Long-term stability test of the u7 catalyst with and without H2S. Gas composition: CH4 (v% = 2%), O2 (v% = 8%), H2S (v% = 0.02%), and N2 (v% = 89.8%); Flow rate: 200 mL/min; and GHSV = 60000 mL/(g·h).
Catalysts 09 00410 g011
Figure 12. Hydrogen sulfide temperature-programmed desorption (H2S-TPD)profiles of the u7 catalyst and freshly prepared PdO/SiO2 catalyst.
Figure 12. Hydrogen sulfide temperature-programmed desorption (H2S-TPD)profiles of the u7 catalyst and freshly prepared PdO/SiO2 catalyst.
Catalysts 09 00410 g012
Scheme 1. Schematic diagram of methane catalytic oxidation on the surface of Pd4S/SiO2 and PdO/SiO2 in presence of H2S.
Scheme 1. Schematic diagram of methane catalytic oxidation on the surface of Pd4S/SiO2 and PdO/SiO2 in presence of H2S.
Catalysts 09 00410 sch001
Table 1. Average particles size of freshly prepared catalysts sulfurized with different H2S concentration.
Table 1. Average particles size of freshly prepared catalysts sulfurized with different H2S concentration.
CatalystFWHMSize (nm)
Pd0.76112.2
f0.20.80311.5
f11.3286.9
f31.2976.7
f53.3292.6
f73.2622.6
f102.1634.0
FWHM: full-width at half-maximum
Table 2. Palladium species content of freshly prepared catalysts with different H2S concentration a.
Table 2. Palladium species content of freshly prepared catalysts with different H2S concentration a.
CatalystComposition Ratio of Palladium Species (%)
Pd0Pd4S
f0.268.931.1
f168.631.4
f361.638.4
f533.466.6
f732.167.9
f1057.942.1
a: Calculated by Pd fitted peak area by XPS.
Table 3. Palladium species content of used catalysts with different H2S concentration a.
Table 3. Palladium species content of used catalysts with different H2S concentration a.
CatalystComposition Ratio of Palladium Species (%)
PdPdOPd4S
u0.25.359.535.2
u14.060.435.6
u35.954.239.9
u52.740.456.9
u73.136.660.3
u103.648.947.5
a: Calculated by Pd fitted peak area by XPS.

Share and Cite

MDPI and ACS Style

Ma, L.; Yuan, S.; Jiang, T.; Zhu, X.; Lu, C.; Li, X. Pd4S/SiO2: A Sulfur-Tolerant Palladium Catalyst for Catalytic Complete Oxidation of Methane. Catalysts 2019, 9, 410. https://doi.org/10.3390/catal9050410

AMA Style

Ma L, Yuan S, Jiang T, Zhu X, Lu C, Li X. Pd4S/SiO2: A Sulfur-Tolerant Palladium Catalyst for Catalytic Complete Oxidation of Methane. Catalysts. 2019; 9(5):410. https://doi.org/10.3390/catal9050410

Chicago/Turabian Style

Ma, Lei, Shiyan Yuan, Taotao Jiang, Xiangming Zhu, Chunshan Lu, and Xiaonian Li. 2019. "Pd4S/SiO2: A Sulfur-Tolerant Palladium Catalyst for Catalytic Complete Oxidation of Methane" Catalysts 9, no. 5: 410. https://doi.org/10.3390/catal9050410

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