Next Article in Journal
Theoretical Investigation of the Deactivation of Ni Supported Catalysts for the Catalytic Deoxygenation of Palm Oil for Green Diesel Production
Next Article in Special Issue
One-Pot Synthesis of Nano CuO-ZnO Modified Hydrochar Derived from Chitosan and Starch for the H2S Conversion
Previous Article in Journal
Transformation of Resinous Components of the Ashalcha Field Oil during Catalytic Aquathermolysis in the Presence of a Cobalt-Containing Catalyst Precursor
Previous Article in Special Issue
Application of Response Surface Methodology for Preparation of ZnAC2/CAC Adsorbents for Hydrogen Sulfide (H2S) Capture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Selective Catalytic Oxidation of Lean-H2S Gas Stream to Elemental Sulfur at Lower Temperature

Department of Industrial Engineering, University of Salerno, Via Giovanni Paolo II, 132, 84084 Fisciano, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(6), 746; https://doi.org/10.3390/catal11060746
Submission received: 28 May 2021 / Revised: 16 June 2021 / Accepted: 16 June 2021 / Published: 18 June 2021
(This article belongs to the Special Issue Catalysts and Processes for H2S Conversion to Sulfur)

Abstract

:
Ceria-supported vanadium catalysts were studied for H2S removal via partial and selective oxidation reactions at low temperature. The catalysts were characterized by N2 adsorption at 77 K, Raman spectroscopy, X-ray diffraction techniques, and X-ray fluorescence analysis. X-ray diffraction and Raman analysis showed a good dispersion of the V-species on the support. A preliminary screening of these samples was performed at fixed temperature (T = 327 °C) and H2S inlet concentration (10 vol%) in order to study the catalytic performance in terms of H2S conversion and SO2 selectivity. For the catalyst that exhibited the higher removal efficiency of H2S (92%) together with a lower SO2 selectivity (4%), the influence of temperature (307–370 °C), contact time (0.6–1 s), and H2S inlet concentration (6–15 vol%) was investigated.

1. Introduction

Hydrogen sulfide (H2S) is a common gas pollutant, which is harmful to human health with deleterious effects on many industrial catalysts, and represents the main source of acid rains when it is oxidized to sulfur dioxide (SO2) [1]. Many attempts have been focused on H2S removal from gaseous streams due to the worldwide increase in restrictive emission standards. Today, H2S-removal-based processes include wet scrubbing [2], biological methods [3], adsorption [4], and selective catalytic oxidation [5]. Among these purification processes, selective catalytic oxidation seems to be very promising for lean-H2S gas streams, where the concentration of hydrogen sulfide is in the range 0.1–10 vol%.
Typically, lean-H2S gas is characteristic of tail gas treating (<5 wt% H2S), crude petroleum (0.3–0.8 wt% H2S), and natural gas streams (0.03–0.3 wt% H2S), although in this last case the H2S can also reach 30 wt% [6].
The selective catalytic oxidation of H2S into elemental sulfur is one of the treatment methods employed for the removal of H2S from the Claus process tail gas [7,8]. This reaction can be performed above or below the sulfur dew point (180 °C) and the processes used are super-Claus, doxosulfreen (Elf-Lurgi), and the mobil direct oxidation process (MODOP) [9]. The super-Claus process, developed in 1985, is continuously being improved and allows achievement of a desulfurization efficiency of ~99.5% at 240 °C in the presence of iron- and chromium-based catalysts supported on alumina or silica [10]. In MODOP, the direct oxidation of H2S into elemental sulfur occurs on a TiO2-based catalyst that deactivates in the presence of water [11]. In the super-Claus process, H2S is oxidized without removing water from the tail gas. Metal-oxide-based catalysts, such as Al2O3, TiO2, V2O5, Mn2O3, Fe2O3, and CuO are the most used and investigated for H2S-selective catalytic oxidation [12]. Indeed, vanadium oxides have been investigated as active phases for H2S selective oxidation and are used as bulk V2O5 [13], mixed with other metals [14], or supported over commercial [15] and mesoporous materials [16].
In our previous works, vanadium-based catalysts supported on different metal oxides (CeO2, TiO2, and CuFe2O4) were investigated for H2S removal from biogas by partial and selective oxidation reactions in the temperature range 50–250 °C [17]. The optimization of the V2O5 loading (2.55–50 wt%) was performed on the CeO2 support at the temperature of 150 °C [18]. The 20 wt% V2O5/CeO2 catalyst showed the best catalytic performance in terms of H2S conversion (99%) and sulfur selectivity (99%) at 150 °C, by feeding a very diluted stream containing only 500 ppm of H2S [19]. Structured catalysts starting from a cordierite carrier in the form of a monolith honeycomb were also prepared, characterized, and tested at low temperature and evidenced high activity and very low SO2 selectivity [20,21].
Based on these obtained promising results, in this study, vanadium-based catalysts supported on ceria were prepared, characterized, and tested in the presence of a lean-H2S gas stream containing a H2S concentration higher than 5 vol%, which is a typical concentration of the Claus process tail gas. A preliminary screening of the catalysts with different vanadium loadings was carried out at 327 °C, in order to identify the catalyst formulation able to maximize the H2S conversion and depress the SO2 formation in the presence of 10 vol% of H2S. The effect of the main operating parameters, such as temperature, contact time, and H2S inlet concentration, was also investigated.

2. Results and Discussion

2.1. Catalytic Activity Test

First of all, the reaction system was studied in the presence of 10 vol% of H2S at the temperature of 327 °C without the catalyst (Figure 1).
Figure 1 shows the behavior of H2S, H2O, and SO2 during 1 h of time on stream.
After the first 5 min, the feed stream was sent to the reactor and the formation of SO2 and water could be observed. The sulfur formation was not detectable because of the removal by the gaseous stream in the sulfur trap. The final H2S conversion was 26%, while the SO2 selectivity was high enough (~39%). The SO3 formation (m/z = 80) was not observed either for the test in the absence of a catalyst or for all the catalytic tests. In Table 1, the values obtained by the test carried out without the catalyst were compared with the ones expected by the thermodynamic equilibrium and with the experimental data achieved with 20 V-CeO2 catalyst.
It is evident that the reaction system without the catalyst is very far from the equilibrium conditions; in fact, the expected H2S conversion and SO2 concentration would be, respectively, 90% and 0.5 vol%. Conversely, the catalytic performance of the 20 V-CeO2 sample is very close to that expected from the equilibrium, confirming the key role of the catalyst for maximizing the H2S conversion and inhibiting the SO2 formation.
The screening of the vanadium-based catalysts was performed at 327 °C and the catalytic activity of the V-CeO2 samples was also compared with the support (CeO2) and with the bulk V2O5. For each sample, the catalytic performance under steady-state conditions is reported in Figure 2.
The best catalytic performance can be observed for the catalysts having a V2O5 loading of 2.55 and 20 wt%, for which the H2S conversion and SO2 selectivity values are very similar. Although the lowest SO2 selectivity (2.2%) was observed for the 50 V-CeO2 sample, it unfortunately showed the lowest H2S conversion (83%).
The influence of the temperature was investigated for the 20 V-CeO2 catalyst and the experimental data for H2S conversion and SO2 selectivity were compared with the equilibrium data (red and blue lines, respectively) (Figure 3).
As it is possible to observe from Figure 3, the H2S conversion is very close to the equilibrium values (red line) while the SO2 selectivity is, in the overall investigated temperature range, slightly below the equilibrium calculation (blue line), evidencing that the catalyst is able to inhibit the SO2 formation. The effect of the H2S concentration, in the range 6–15 vol%, was then evaluated at the temperature of 327 °C (Figure 4).
The highest value of H2S conversion and the lowest SO2 selectivity were observed when the H2S inlet concentration was 10 vol%. In the presence of a feed stream more concentrated in H2S (15 vol%), the conversion was drastically reduced to 75% and the SO2 concentration was about 1.5 vol%; in this case, the selectivity increase is of one magnitude order (14%) with respect to the other obtained values. In Table 2, the equilibrium data are compared with those obtained experimentally at different H2S inlet concentrations.
Based on the data listed in Table 2, it is possible to see that the reaction system deviates from the equilibrium values especially in presence of 15 vol% of H2S.
The influence of the contact time on the catalytic performance is reported in Figure 5. For comparison, the equilibrium data for both H2S conversion and SO2 selectivity at the temperature of 327 °C are also shown.
The catalytic performance resulted in little affected from the variation of the contact time. In particular, it is noteworthy to evidence that the H2S conversion is quite close to the equilibrium values, while the SO2 selectivity is in all cases below the equilibrium value.

2.2. Catalyst Characterization

The nominal and measured vanadium oxide content of the catalysts before the activation step is reported in Table 3.
The results reported evidence that the nominal V2O5 loading is very close to the measured loading.
The specific surface areas of the fresh and used catalysts are reported in Table 4.
The lowest SSA was observed for the sample that was not supported (V2O5). In particular, the value of bulk V2O5 (8 m2/g) decreased more than 50% after the catalytic test. For the fresh V-CeO2 catalysts, the values of SSA were slightly lower than the CeO2 support (~30 m2/g). After the catalytic activity tests, the SSA decrease was likely due to the sulfur deposition on the catalyst surface. This aspect was more evident for the 20 V-CeO2 sample (SSA = 4 m2/g) and was confirmed by XRD and Raman characterizations.
Raman spectra of the support and fresh catalysts are shown in Figure 6. The Raman spectrum for pure CeO2 shows the main band at 460 cm−1, ascribable to ceria in the typical cubic crystal structure of fluorite-type cerium oxide [22,23]. The 2.55 V-CeO2 sample shows that such Raman band slightly shifted to 465 cm−1, while in the case of the catalysts with the highest V loading this band shifted up to 454 cm−1. A more detailed discussion of these results is reported in the Supplementary Materials (Figure S1).
The XRD spectra of CeO2 and the fresh catalysts are shown in Figure 7. All the catalysts exhibit the characteristic peaks of CeO2 at 28.3°, 32.8°, 47.3°, 56.1°, 58°, and 69°, corresponding to diffraction planes indexed as (1 1 1), (2 0 0), (2 2 0), (3 1 1), (2 2 2), and (4 0 0), respectively [24]. These patterns are ascribable to the typical cubic crystal structure of fluorite-type cerium oxide [25]. No additional reflections attributable to V2O5 are detectable, evidencing that the sulfuration of the catalysts completely occurred [26]. Furthermore, there were no peaks detected that related to typical vanadium sulfides (VS2, VS4, V2S3, V3S4) that might have formed following the sulfuration treatment [27].
In Figure 8, the Raman spectra of the fresh samples (CeO2 and V-CeO2 catalysts) are compared with the used catalysts. The used CeO2, equally to the fresh one, has the characteristic Raman peak perfectly centered at 460 cm−1 (Figure 8a) [22,23]. A slight shift of this Raman band up to 465 cm−1, 457 cm−1, and 454 cm−1 (Figure 8b–d) is detectable for 2.55 V-CeO2, 20 V-CeO2, and 50 V-CeO2 fresh catalysts, respectively, as already previously observed (Figure 6). A detailed discussion of the Raman results is reported in the Supplementary Materials (Figure S2).
Furthermore, the absence of any characteristic bands of the vibrational modes of crystalline V2O5 [28] and V = O stretching vibration ascribable to monovanadate species (VO43−) denotes that the sulfuration of the catalysts occurred completely [16]. The Raman spectrum of the bulk V2O5 after the catalytic test is reported in Figure 9.
The Raman bands at 140, 192, 282, 405, 688, and 993 cm−1 are characteristic of the vanadium sulfide in VS2 form, as reported in the literature [29]. In particular, all the signals correspond to the rocking combination and stretching vibrations of V–S bonds or their combination [30]. Moreover, no bands related to the formation of vanadyl sulfate (984 cm−1 and 1060 cm−1) were observed [31].
In Figure 10, the XRD patterns of the fresh samples are compared with the used ones. There are no differences between the XRD spectra of the fresh/used bulk CeO2; for the used sample less intensity of the peaks is observed, which is likely due to the sulfur deposition (Figure 10a). For the used 2.55 V-CeO2 catalyst, in addition to the characteristic peaks of the CeO2 fresh sample, a signal is visible at 2θ = 23° due to the sulfur formation [32], as also confirmed from Raman analysis (Figure 10b). The spectra of the used 20 V-CeO2 catalyst (Figure 10c) are different, where other peaks attributable to the sulfur are observable at 2θ = 23°, 24°, 26°, 27°, and 28° [32]. For the 50 V-CeO2 catalyst, the XRD spectrum of the fresh sample is perfectly stackable with that of the used sample (Figure 10d) because all the peaks are ascribable only to the CeO2 support.
The average crystallite size of ceria for the different catalysts, calculated with the Scherrer equation, are listed in Table 5.
As it is possible to observe from Table 5, the increase of the V-loading for the different catalysts has involved an increase in the crystallite size of the CeO2, as reported in the literature for supported vanadium catalysts [25]. The average crystallite size of bulk CeO2 before the catalytic tests was 13 nm; it increased to 24 nm for the catalyst having the highest V-content (50 V-CeO2). Relatively to the catalysts, there is a negligible variation of the ceria average crystallite size between fresh and used samples.
The only significant variation between fresh and used samples was obtained for the support; the greater segregation of the CeO2 after the catalytic activity tests involved the increase of the crystallite dimension (21 nm). The segregation of the CeO2 crystallite may be due to the high SO2 formation observed on the support in the absence of the active phase; in fact, among the catalysts, the highest value of SO2 selectivity (~10%) was obtained for the CeO2 at 327 °C as previously reported in Figure 2. The reaction temperature could favor the formation of sulfate species and also the oxygen in the ceria lattice could facilitate the CeO2 sulfuration [33]; therefore, the reaction between CeO2 and SO2 could occur, leading to the formation of cerium sulfate Ce(SO4)2, which is stable at high temperature and decomposes between 722 and 843 °C to CeO2 [34].

3. Materials and Methods

3.1. Catalyst Preparation and Characterization

The preparation of vanadium-based catalysts supported on ceria with different loading of active phase (2.55, 20, and 50 wt% V2O5 nominal loading) was described in detail in our previous work [19]. All the reactants were provided by Sigma Aldrich. After the calcination, the sulfuration procedure was carried out in a quartz reactor containing the catalyst to be sulfurized. In particular, the activation step was realized by feeding a gaseous stream containing N2 and H2S at 20 vol%, by increasing the temperature from ambient temperature up to 200 °C with a heating rate of 10 °C/min for 1 h. Finally, the catalysts were reduced to the size 38–180 µm. For simplicity, the catalysts are named in the paper as follows: 2.55 V-CeO2, 20 V-CeO2, 50 V-CeO2, where “2.55” means the nominal V loading (wt%) expressed as V2O5. The sulfurized samples before the testing are named “fresh”, while they are named “used” after the catalytic activity test.
The catalysts were characterized by nitrogen adsorption at 77 K, Raman spectroscopy and X-ray diffraction. The specific surface area was evaluated with a Costech Sorptometer 1040 (Costech International, Firenze, Italy) by using N2 and He, respectively, as adsorptive and carrier gas. The powder catalysts were treated at 150 °C for 30 min in a He flow prior to testing. A BET method multipoint analysis based on N2 adsorption/desorption isotherms at 77 K was used to evaluate the specific surface area of the fresh and used catalysts. X-ray diffraction (XRD) was performed using a Brucker D2 Phaser (Germany) using CuKa radiation (λ = 1.5401 A°). Laser Raman spectra of the catalysts were obtained in air with a Dispersive MicroRaman (Invia, Renishaw, Italy), equipped with a 514 nm diode-laser, in the range of 100–2000 cm−1 Raman shift. The V-content of the fresh catalysts (expressed as V2O5 wt%) was evaluated by X-ray fluorescence (XRF) spectra by using an ARL QUANT’X EDXRF spectrometer (ThermoFisher Scientific, Italy).

3.2. Experimental Apparatus

The catalytic activity tests were performed in the laboratory plant schematized in Figure 11.
The laboratory plant is made of three sections: feed, reaction, and analysis sections. The feed stream containing H2S, O2, and N2 is sent by a three-way valve to the reactor, or in bypass position to the analyzer to verify the composition. All gases came from SOL S.p.A with a purity degree of 99.999% for N2, O2, and SO2, and 99.5% for H2S.
The reaction system comprises a furnace, a reactor, and a sulfur abatement trap. The quartz-made reactor, consisting of a tube of 300 mm length and an internal diameter of 19 mm, is housed in a vertical furnace heated with silicon carbide (SiC)-based resistances. At the bottom of the reactor are a reactant inlet and a thermocouple sheet concentric to the reactor. The catalytic bed is placed in the isothermal zone of the reactor and the temperature is measured continuously by a K-type thermocouple. In the head of the reactor is welded a trap for the sulfur abatement, which is made of an expansion vessel that allows the sulfur to liquefy, involving its separation by the gaseous stream. This trap is maintained at the temperature of 250 °C.
All the lines downstream of the reactor were heated at the temperature of 170 °C to avoid sulfur solidification and possible clogging of the mass spectrometer capillary and to maintain the water in the gas phase for the analysis. The analysis of the gaseous stream (H2S, O2, N2, SO2, SO3) was performed with the mass spectrometer quadrupole (Hiden HPR-20) (Warrington, United Kingdom).
Finally, the abatement of unconverted H2S was realized by adsorption on activated carbons loaded in a special vessel having a capacity of 10 Lt. Furthermore, the entire apparatus plant was housed under the hood and isolated from the external environment in order to avoid gas leakage.
The operating conditions of the catalytic activity tests are listed in Table 6.
H2S conversion (x H2S) and the SO2 selectivity (s SO2) were calculated by using the following relationship (Equations (1) and (2)), by considering the gas phase volume change to be negligible:
x H2S, % = ((H2SIN − H2SOUT)/H2SIN)·100
s SO2, % = (SO2OUT/(H2SIN − H2SOUT))·100
For the equilibrium calculation, the GasEq program was used, software (0.7.0.9 version, Chris Morley) based on the minimization of Gibbs free energy, which is able to calculate the equilibrium product composition of an ideal gaseous mixture when there are a lot of simultaneous reactions. The thermodynamic analysis was carried out considering the following chemical species that could be present at equilibrium: H2S, O2, SO2, S2, S6, S8, H2O, and nitrogen.

Calibration Procedure

The calibration procedure is required in order to measure the concentration of all the species that could be in the gas stream for analysis and, for this reason, it must be performed prior to carrying out experimental tests. However, it could be necessary to repeat the calibration every time the process conditions are changed (e.g., after the replacement of the capillary, the filaments, change of pressure chamber value) or when the signal seems to be affected by derivative effects. The measurements could be affected by interference due to the presence of ions of different molecules having the same m/z ratio. Each molecule has a matrix of interference, which defines the “weight” of the disturbance of other molecules on the partial pressure of the molecule in the phase of calibration. The partial pressure obtained, net of the relative interference, must be corrected by a response factor, thus returning the actual partial pressure of each molecule in the stream analyzed. At this point, it is possible to calculate the correct concentration of each component. The calibration procedure is characterized by different steps:
(1)
Report in a table the partial pressure of the all-mass fragment for each concentration of the component to calibrate;
(2)
Construct the matrix of interference and the response factors table relatively for the component you are calibrating;
(3)
Calculate the concentrations of the component calibrated by considering the relative interference of other species on the component to calibrate and correcting the measure by its response factor.
After the calibration, which is carried out in a by-pass position, the feed stream can be sent to the reactor. The reactor, before each test, is purged with nitrogen to avoid humidity and/or impurity and is heated up to the reaction temperature at which the feed stream is sent.
Similarly, at the end of the activity test, the reactor is cooled down with nitrogen to room temperature.

4. Conclusions

The H2S selective oxidation reaction to sulfur and water was investigated over vanadium-sulfide-based catalysts supported on CeO2. The catalysts were prepared with different vanadium loading and were characterized before and after the catalytic tests with different techniques. X-ray diffraction and Raman analysis showed a good dispersion of the V-species on the support because the V-sulfide presence was not detected on the different catalysts. The only vanadium sulfide in VS2 form was observed for the bulk V2O5 after the catalytic tests. Furthermore, the presence of the sulfur was observed especially over the used catalysts at lower V-loading by Raman and SSA analysis.
From the preliminary screening of the catalysts performed at 327 °C, the higher catalytic activity was observed over the 2.55 V-CeO2 and 20 V-CeO2 catalysts, with H2S conversion, respectively, of 90% and 92%, and SO2 selectivity of ~4%. No SO3 formation and catalyst deactivation phenomena by the sulfur deposition were observed. The effect of the temperature, contact time, and H2S inlet concentration was studied over 20 V-CeO2 catalysts. By increasing the H2S inlet concentration (up to 15 vol%), the conversion decreased from 86% to 75% with an SO2 concentration of about 1.5 vol%. The effect of the contact time was almost negligible on the H2S conversion and SO2 selectivity, while the temperature had a significant influence. In the range of temperatures investigated (300–370 °C), the H2S conversion was very close to the equilibrium values while the SO2 selectivity was below the equilibrium calculation, evidencing that the catalyst is effectively able to inhibit SO2 formation.
Based on the obtained results, the ceria-supported vanadium catalysts could be considered good candidates to carry out the selective oxidation of H2S to sulfur by an H2S-lean gas stream (e.g., natural gas, Claus process tail gas) at very low temperature.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11060746/s1: Figure S1: Raman Spectra of CeO2, V2O5 and 2.55, 20, 50 V-CeO2 fresh catalysts; Figure S2: Raman Spectra of fresh and used catalysts CeO2 (a), 2.55 V-CeO2 (b), 20 V-CeO2 (c), 50 V-CeO2 (d).

Author Contributions

Conceptualization, D.B. and V.V.; methodology, D.B.; software, D.B.; validation, D.B. and V.V.; formal analysis, V.P.; investigation, D.B.; resources, V.V. and V.P.; data curation, V.P.; writing—original draft preparation, D.B.; writing—review and editing, V.V.; visualization, V.P. and V.V; supervision, V.P.; project administration, D.B.; funding acquisition, V.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Forzatti, P.; Lietti, L. Catalyst deactivation. Catal. Today 1999, 52, 165–181. [Google Scholar] [CrossRef]
  2. Wang, R. Investigation on a new liquid redox method for H2S removal and sulfur recovery with heteropoly compound. Sep. Purif. Technol. 2003, 31, 111–121. [Google Scholar] [CrossRef]
  3. Duan, H.; Yan, R.; Koe, L.C.; Wang, X. Combined effect of adsorption and biodegradation of biological activated carbon on H2S biotrickling filtration. Chemosphere 2007, 66, 1684–1691. [Google Scholar] [CrossRef]
  4. Seredych, M.; Bandosz, T.J. Sewage sludge as a single precursor for development of composite adsorbents/catalysts. Chem. Eng. J. 2007, 128, 59–67. [Google Scholar] [CrossRef]
  5. Yasyerli, S.; Dogu, G.; Ar, I.; Dogu, T. Dynamic analysis of removal and selective oxidation of H2S to elemental sulfur over Cu–V and Cu–V–Mo mixed oxides in a fixed bed reactor. Chem. Eng. Sci. 2004, 59, 4001–4009. [Google Scholar] [CrossRef]
  6. Elmawgoud, H.A.; Elshiekh, M.; Abdelkreem, M.; Khalil, S.A.; Alsabagh, A.M. Optimization of petroleum crude oil treatment using hydrogen sulfide scavenger. Egypt. J. Pet. 2019, 28, 161–164. [Google Scholar] [CrossRef]
  7. Pi, J.H.; Lee, D.H.; Lee, J.D.; Jun, J.H.; Park, N.K.; Ryu, S.O.; Lee, T.J. The study on the selective oxidation of H2S over the mixture zeolite NaX-WO3 catalysts. Korean J. Chem. Eng. 2004, 21, 126–131. [Google Scholar] [CrossRef]
  8. Lee, J.D.; Han, G.B.; Park, N.K.; Ryu, S.O.; Lee, T.J. The selective oxidation of H2S on V2O5/zeolite-X catalysts in an IGCC system. J. Ind. Eng. Chem. 2006, 12, 80–85. [Google Scholar]
  9. Wiheeb, A.D.; Shamsudin, I.K.; Ahmad, M.A.; Murat, N.M.; Kim, J.; Othman, M.R. Present technologies for hydrogen sulfide removal from gaseous mixtures. Rev. Chem. Eng. 2013, 29, 449–470. [Google Scholar] [CrossRef]
  10. Keller, N.; Pham-Huu, C.; Crouzet, C.; Ledoux, M.J.; Savin-Poncet, S.; Nougayrede, J.B.; Bousquet, J. Direct oxidation of H2S into S: New catalysts and processes based on SiC support. Catal. Today 1999, 53, 535–542. [Google Scholar] [CrossRef]
  11. Zhang, X.; Tang, Y.; Qu, S.; Da, J.; Hao, Z. H2S-selective catalytic oxidation: Catalysts and processes. ACS Catal. 2015, 5, 1053–1067. [Google Scholar] [CrossRef]
  12. Davydov, A.A.; Marshneva, V.I.; Shepotko, M.L. The comparison study of the catalytic activity. Mechanism of the interactions between H2S and SO2 on some oxides. Appl. Catal. A 2003, 244, 93–100. [Google Scholar] [CrossRef]
  13. Li, K.T.; Hyang, M.Y.; Cheng, W.D. Vanadium-based mixed-oxide catalysts for selective oxidation of hydrogen sulfide to sulfur. Ind. Eng. Chem. Res. 1996, 35, 621–626. [Google Scholar] [CrossRef]
  14. Park, D.W.; Byung, B.H.; Ju, W.D.; Kim, M.I.; Kim, K.H.; Woo, H.C. Selective oxidation of hydrogen sulfide containing excess water and ammonia over Bi-V-Sb-O catalysts. Korean J. Chem. Eng. 2005, 22, 190–195. [Google Scholar] [CrossRef]
  15. Kalinkin, P.; Kovalenko, O.; Lapina, O.; Khabibulin, D.; Kundo, N. Kinetic peculiarities in the low-temperature oxidation of H2S over vanadium catalysts. J. Mol. Catal. A Chem. 2002, 178, 173–180. [Google Scholar] [CrossRef]
  16. Soriano, M.D.; Jimenez-Jimenez, J.; Concepcion, P.; Jimenez- Lopez, A.; Rodrıguez-Castellon, E.; Lopez Nieto, J.M. Selective oxidation of H2S to sulfur over vanadia supported on mesoporous zirconium phosphate heterostructure. Appl. Catal. B Environ. 2009, 92, 271–279. [Google Scholar] [CrossRef]
  17. Palma, V.; Barba, D.; Ciambelli, P. Screening of catalysts for H2S abatement from biogas to feed molten carbonate fuel cells. Int. J. Hydrogen Energy 2013, 38, 328–335. [Google Scholar] [CrossRef]
  18. Palma, V.; Barba, D. Low temperature catalytic oxidation of H2S over V2O5/CeO2 catalysts. Int. J. Hydrogen Energy 2014, 39, 21524–21530. [Google Scholar] [CrossRef]
  19. Palma, V.; Barba, D. Vanadium-ceria catalysts for H2S abatement from biogas to feed to MCFC. Int. J. Hydrogen Energy 2017, 42, 1891–1898. [Google Scholar] [CrossRef]
  20. Palma, V.; Barba, D.; Gerardi, V. Honeycomb-structured catalysts for the selective partial oxidation of H2S. J. Clean Prod. 2016, 111, 69–75. [Google Scholar] [CrossRef]
  21. Palma, V.; Barba, D. Honeycomb V2O5-CeO2 catalysts for H2S abatement from biogas by direct selective oxidation to sulfur at low temperature. Chem. Eng. Trans. 2015, 43, 1957–1962. [Google Scholar] [CrossRef]
  22. Gu, X.; Ge, J. Structural, redox and acid–base properties of V2O5/CeO2 catalyst. Thermochim. Acta 2006, 451, 84–93. [Google Scholar] [CrossRef]
  23. Escribano, V.S.; López, E.F.; Panizza, M.; Resini, C.; Amores, J.M.G.; Busca, G. Characterization of cubic ceria–zirconia powders by X-ray diffraction and vibrational and electronic spectroscopy. Solid State Sci. 2003, 5, 1369–1376. [Google Scholar] [CrossRef]
  24. Sun, C.; Li, H.; Zhang, H.; Wang, Z.; Chen, L. Controlled synthesis of CeO2 nanorods by a solvothermal method. Nanotechnology 2005, 16, 1454–1463. [Google Scholar] [CrossRef]
  25. Radhika, T.; Sugunan, S. Structural and catalytic investigation of vanadia supported on ceria promoted with high surface area rice husk silica. J. Mol. Catal. A Chem. 2006, 250, 169–176. [Google Scholar] [CrossRef] [Green Version]
  26. Singh, B.; Gupta, M.K.; Mishra, S.K.; Mittal, R.; Sastry, P.U.; Rolsc, S.; Lal Chaplot, S. Anomalous lattice behavior of vanadium pentaoxide (V2O5): X-ray diffraction, inelastic neutron scattering and ab initio lattice dynamics. Phys. Chem. Chem. Phys. 2017, 19, 17967. [Google Scholar] [CrossRef] [Green Version]
  27. Liu, Y.Y.; Xu, L.; Guo, X.T.; Lv, T.T.; Pang, H. Vanadium sulfide based materials: Synthesis, energy storage and conversion. J. Mater. Chem. 2020, 8, 20781–20802. [Google Scholar] [CrossRef]
  28. Holgrado, J.P.; Soriano, M.D.; Jimenez, J. Operando XAS and Raman study on the structure of a supported vanadium 409 oxide catalyst during the oxidation of H2S to sulfur. Appl. Catal. B 2010, 92, 271–279, 410. [Google Scholar] [CrossRef]
  29. Huang, L.; Zhu, W.; Zhang, W.; Chen, K.; Wang, J.; Wang, R.; Yang, Q.; Hu, N.; Suo, Y.; Wang, J. Layered vanadium (IV) 414 disulfide nanosheets as a peroxidase-like nanozyme for colorimetric detection of glucose. Microchim. Acta 2018, 185, 415. [Google Scholar] [CrossRef]
  30. Qu, Y.; Shao, M.; Shao, Y.; Yang, M.; Xu, J.; Kwok, C.T.; Shi, X.; Lu, Z.; Pan, H. Ultra-high electrocatalytic activity of VS2 nanoflowers for efficient hydrogen evolution reaction. J. Mater. Chem. A 2017, 5, 15080–15086. [Google Scholar] [CrossRef]
  31. Evans, J.C. The vibrational spectra and structure of the vanadyl ion in aqueous solution. Inorg. Chem. 1963, 2, 372–375. [Google Scholar] [CrossRef]
  32. Xu, J.; Su, D.; Zhang, W.; Bao, W.; Wang, G. A nitrogen–sulfur co-doped porous graphene matrix as a sulfur immobilizer for high performance lithium–sulfur batteries. J. Mater. Chem. A 2016, 4, 17381–17393. [Google Scholar] [CrossRef] [Green Version]
  33. Smirnov, M.Y.; Kalinkin, A.V.; Pashis, A.V.; Sorokin, A.M. Interaction of Al2O3 and CeO2 Surfaces with SO2 and SO2 + O2 studied by X-ray photoelectron spectroscopy. J. Phys. Chem. B 2005, 109, 11712–11719. [Google Scholar] [CrossRef] [PubMed]
  34. Sharma, I.B.; Singh, V.; Lakhanpal, M. Study of thermal decomposition of ammonium cerium sulphate. J. Therm. 1992, 38, 1345–1355. [Google Scholar] [CrossRef]
Figure 1. Activity test without catalyst (T = 327 °C, H2S = 10 vol%, residence time = 0.6 s).
Figure 1. Activity test without catalyst (T = 327 °C, H2S = 10 vol%, residence time = 0.6 s).
Catalysts 11 00746 g001
Figure 2. Catalytic performance of the different catalysts under steady-state conditions (T = 327 °C, H2S = 10 vol%, contact time = 0.6 s).
Figure 2. Catalytic performance of the different catalysts under steady-state conditions (T = 327 °C, H2S = 10 vol%, contact time = 0.6 s).
Catalysts 11 00746 g002
Figure 3. Temperature effect over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (H2S = 10 vol%, contact time = 0.6 s).
Figure 3. Temperature effect over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (H2S = 10 vol%, contact time = 0.6 s).
Catalysts 11 00746 g003
Figure 4. Influence of the H2S inlet concentration over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (O2/H2S = 0.5, T = 327 °C, contact time = 0.6 s).
Figure 4. Influence of the H2S inlet concentration over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (O2/H2S = 0.5, T = 327 °C, contact time = 0.6 s).
Catalysts 11 00746 g004
Figure 5. Effect of the contact time over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (T = 327 °C, H2S = 10 vol%).
Figure 5. Effect of the contact time over 20 V-CeO2 catalyst on the H2S conversion and SO2 selectivity (T = 327 °C, H2S = 10 vol%).
Catalysts 11 00746 g005
Figure 6. Raman spectra of CeO2, V2O5, and 2.55, 20, 50 V-CeO2 fresh catalysts.
Figure 6. Raman spectra of CeO2, V2O5, and 2.55, 20, 50 V-CeO2 fresh catalysts.
Catalysts 11 00746 g006
Figure 7. XRD spectra of CeO2 and 2.55, 20, 50 V-CeO2 fresh catalysts.
Figure 7. XRD spectra of CeO2 and 2.55, 20, 50 V-CeO2 fresh catalysts.
Catalysts 11 00746 g007
Figure 8. Raman spectra of fresh and used catalysts: CeO2 (a), 2.55 V-CeO2 (b), 20 V-CeO2 (c), 50 V-CeO2 (d).
Figure 8. Raman spectra of fresh and used catalysts: CeO2 (a), 2.55 V-CeO2 (b), 20 V-CeO2 (c), 50 V-CeO2 (d).
Catalysts 11 00746 g008
Figure 9. Raman spectra of V2O5 used.
Figure 9. Raman spectra of V2O5 used.
Catalysts 11 00746 g009
Figure 10. XRD spectra of fresh and used catalysts: CeO2 (a), 2.55 V-CeO2 (b), 20 V-CeO2 (c), 50 V-CeO2 (d).
Figure 10. XRD spectra of fresh and used catalysts: CeO2 (a), 2.55 V-CeO2 (b), 20 V-CeO2 (c), 50 V-CeO2 (d).
Catalysts 11 00746 g010
Figure 11. Scheme of the apparatus plant.
Figure 11. Scheme of the apparatus plant.
Catalysts 11 00746 g011
Table 1. Comparison between non-catalytic system, catalytic system, and equilibrium (T = 327 °C, H2S = 10 vol%).
Table 1. Comparison between non-catalytic system, catalytic system, and equilibrium (T = 327 °C, H2S = 10 vol%).
No Catalyst20 V-CeO2Equilibrium
H2S Conversion, %26 (±1.5)92 (±1.5)90
SO2 Selectivity, %38.5 (±2)4 (±2)6
SO2, vol%10.40.5
Table 2. Equilibrium and experimental data by varying the H2S inlet concentration (T = 327 °C, contact time = 0.6 s).
Table 2. Equilibrium and experimental data by varying the H2S inlet concentration (T = 327 °C, contact time = 0.6 s).
H2SIN, vol%xH2S, %xH2S Eq., %S SO2, %sSO2 Eq., %
688 (±1.5)894.4 (±2)4
1092 (±1.5)904 (±2)6
1575 (±1.5)9014 (±2)9
Table 3. Theoretical and measured vanadium content of the catalysts before the sulfuration.
Table 3. Theoretical and measured vanadium content of the catalysts before the sulfuration.
SampleV2O5 Nominal wt%% V2O5 Measured wt%
2.55 V-CeO22.552.7
20 V-CeO22022
50 V-CeO25051
Table 4. Specific surface area (SSA, m2/g) of the fresh and used catalysts.
Table 4. Specific surface area (SSA, m2/g) of the fresh and used catalysts.
SampleCeO2V2O52.55 V-CeO220 V-CeO250 V-CeO2
Fresh298252220
Used17217414
Table 5. Average crystallite size (<L>, nm) of the fresh and used catalysts.
Table 5. Average crystallite size (<L>, nm) of the fresh and used catalysts.
SampleFreshUsed
CeO21321
2.55 V-CeO21619
20 V-CeO21819
50 V-CeO22425
Table 6. Operating conditions.
Table 6. Operating conditions.
Operating Conditions
Temperature200–367 °C
Contact Time0.6–1 s
Catalyst Volume3 cm3
Total Flow Rate180–300 Ncc·min−1
GHSV3600–6000 h−1
H2SIN concentration6–15 vol%
O2/H2S0.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Barba, D.; Vaiano, V.; Palma, V. Selective Catalytic Oxidation of Lean-H2S Gas Stream to Elemental Sulfur at Lower Temperature. Catalysts 2021, 11, 746. https://doi.org/10.3390/catal11060746

AMA Style

Barba D, Vaiano V, Palma V. Selective Catalytic Oxidation of Lean-H2S Gas Stream to Elemental Sulfur at Lower Temperature. Catalysts. 2021; 11(6):746. https://doi.org/10.3390/catal11060746

Chicago/Turabian Style

Barba, Daniela, Vincenzo Vaiano, and Vincenzo Palma. 2021. "Selective Catalytic Oxidation of Lean-H2S Gas Stream to Elemental Sulfur at Lower Temperature" Catalysts 11, no. 6: 746. https://doi.org/10.3390/catal11060746

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