Hydrogen production over Co-promoted Mo-S water gas shift catalysts supported on ZrO2

https://doi.org/10.1016/j.apcata.2019.117361Get rights and content

Highlights

  • Co addition enhanced the WGS activity of Mo-S/ZrO2 catalysts.

  • The formation of CoMo-S species increased with Co content up to Co/Mo = 0.3.

  • CO conversion was stable during 4 weeks of WGS reaction in 7000 ppm H2S containing feed at 35,000 h−1 GHSV and 350 °C.

  • The optimal atomic Co/Mo ratio at monolayer coverage on ZrO2 resulted in enhanced WGS activity.

  • The active metal loading in the best CoMo-S catalyst was lower than in pure Mo-S/ZrO2 and commercial catalysts.

Abstract

A series of CoMo-S/ZrO2 catalysts were prepared by modifying Co and Mo content using incipient wetness impregnation method, investigated in the water gas shift (WGS) reaction in the presence of H2S and characterized by TEM, XRD, Raman, TPR, and XPS. The supported CoMo-S catalyst corresponding to the CoMo-O monolayer coverage displayed the presence of highly dispersed CoMo-S species that were characterized by the highest extent of sulfidation. The presence of cobalt facilitated the formation of active CoMo-S species that were more H2S-dependent as compared to Mo-S entities. The Co/Mo = 0.3 catalyst at monolayer CoMo-O coverage was the most active and thermally stable, as well as the least H2S-dependent during 4 weeks of WGS reaction among all catalysts investigated.

Introduction

Hydrogen is an environmentally friendly and high energy density (120 MJ/kg) molecule that can be used as a fuel for stationary electric power plants and mobile fuel cells with zero CO2 emissions [[1], [2], [3], [4]]. Most hydrogen (∼96%) is currently produced from fossil fuel feedstocks: 48% from high-temperature steam reforming of natural gas, 30% from partial oxidation of refinery oil, and 18% from coal gasification [5,6]. The major challenges of using fossil fuels for hydrogen production are high CO2 emissions, the uneven distribution, and shortage of fossil fuel reserves [7]. However, hydrogen production from biomass-derived syngas could overcome these challenges, since biomass is a carbon-neutral and abundant sustainable energy resource [[7], [8], [9]].

The water gas shift (WGS) reaction is a key operation in maximizing the production of hydrogen from biomass-derived syngas [7,8]. As the WGS reaction is reversible and mildly exothermic, the WGS activity depends on reaction temperature (CO(g)+H2O(g)CO2(g)+H2(g)[ΔH0=-41.1kJ/mol]) [10,11]. Typical WGS reactors consist of a high-temperature shift (HTS, 350∼450 °C) and a low-temperature shift (LTS, ∼250 °C) stages due to the thermodynamic limitations of this reaction [[10], [11], [12]]. However, it is quite challenging to apply typical WGS catalysts in hydrogen production using biomass-derived syngas, since biomass feedstocks contain a considerable amount of sulfur [13].

The main challenge of using current Fe2O3-Cr2O3 HTS catalysts is the easy deactivation of these catalysts when the H2S concentration in the syngas exceeds 500 ppm [14], while the commercially available Cu–ZnO–Al2O3 (LTS) catalysts deactivate completely in the presence of trace levels of H2S (∼10 ppm) [15]. On the other hand, the molybdenum-based catalysts have gained significant interest as a sulfur-tolerant WGS catalyst in recent years. However, Mo-based catalysts are sulfur-dependent, requiring at least 100 ppm of H2S to display WGS activity [16]. Currently, no catalyst can maintain WGS activity over the entire range of sulfur content encountered in biomass (0 ∼ 15,000 ppm on dry basis) [13]. Moreover, the WGS activity of Mo-based catalysts is relatively low as compared to conventional HTS and LTS WGS catalysts, Fe- and Cu-based [11,17] and needs to be further improved for this application.

As compared to traditional supports, such as alumina, ZrO2 is a promising support for Mo-based WGS catalysts because it can improve their catalytic activity due to weak MoO3-support interactions. For example, alumina-supported Mo catalysts pre-activated above 500 °C interacted strongly with Al2O3 and displayed lower activity in hydrodesulfurization (HDS) of thiophene than these catalysts pre-activated below 400 °C [18]. On the other hand, Shetty et al. reported that the ZrO2-supported Mo catalysts characterized by weak interactions between ZrO2 and Mo species showed high catalytic activity in the hydrodeoxygenation of m-cresol to toluene [19]. The Mo/ZrO2 catalysts showed the greatest extent of MoO3 reduction to MoO2 among the ZrO2-, TiO2-, Al2O3-, SiO2-, and MgO-supported Mo catalysts [20], suggesting that MoO3 interacted weakly with ZrO2 support.

The high density of surface Mo-O sites is traditionally achieved by enhancing their dispersion as a function of MoO3 surface coverage on oxide supports. High hydrodeoxygenation (HDO) of anisole activity was observed over the Mo-O/ZrO2 catalysts containing a monolayer MoO3 surface coverage due to the presence of highly dispersed Mo oxide species [21]. Al2O3-supported Mo-based catalysts at monolayer MoO3 surface coverage showed the highest WGS activity [22,23]. The modified dispersion of Mo-O species on ZrO2 at different MoO3 loadings has been reported by Chary et al. [24]. However, the effects of CoMo-O surface coverage on ZrO2 support have not been widely studied in the WGS reaction.

Several additives were proposed to promote the WGS activity of Mo-based catalysts. Alkali metal promoters have been suggested to improve the WGS activity by enhancing reducibility and sulfidation of Mo [25], while the high mobility of alkali promoters at high reaction temperature had a negative impact on the stability of Mo-based catalysts [26]. Nickel (Ni)-promoted Mo-S catalysts have been reported to show enhanced WGS activity and bimetallic synergistic effect [[27], [28], [29], [30]], whereas the methanation side-reaction and increased carbon deposition are typical disadvantages of Ni-Mo catalysts for hydrogen production. Cobalt was reported as the optimal additive to improve the WGS activity of Mo-based catalysts by increasing the density of active sites [[31], [32], [33], [34], [35]]. CoMo-S/TiO2-Al2O3 catalysts showed higher WGS activity than commercial sour shift catalysts (CoMo-MgO-Al2O3) at H2O/CO = 1 [36,37].

The WGS reaction pathways for the Al2O3-supported Mo-S and Co-promoted Mo-S species have been investigated by DFT calculations and in experimental studies [38,39]. The CoMo-S and Mo-S sites in the CoMo-S/Al2O3 catalysts were distinguished and quantified using CO adsorption followed by in-situ IR spectroscopy (IR/CO) at low temperature [38,40]. Chen et al. reported that the intrinsic activity of CoMo-S sites in the WGS reaction at low temperature (< 300 °C) was higher than that of unpromoted Mo-S sites [38]. Recent DFT calculations over CoMo-S WGS catalysts reported that the overall reaction barrier for the COOH-associated mechanism proposed for the WGS reaction on the S edge of Co-MoS2 was lower than that of the redox mechanism on the S edge of MoS2 [39]. The Co addition can reduce the height of the H2O dissociation energy (H2O → OH + H) on the sulfur edge sites of Co-MoS2 as compared to that of unpromoted Mo-S [39,41], which is one of the rate-determining steps in the WGS reaction.

We investigated two series of ZrO2 supported CoMo-S catalysts. In the first series, we investigated the influence of CoMo-O surface coverage ranging from 0.5 to 4 layers on ZrO2 on the WGS activity. In the second series, the effect of Co/Mo ratios at monolayer surface coverage were investigated. The long-term stability and H2S-dependence of CoMo-S/ZrO2 catalysts as a function of CoMo-O surface coverage and Co/Mo atomic ratios were also investigated employing an H2S containing feed, while both series of CoMo-S/ZrO2 catalysts were characterized by TEM, XRD, Raman, TPR, and XPS.

Section snippets

Catalyst synthesis

Supported CoMo catalysts were prepared by incipient wetness impregnation of the ZrO2 support with a solution of ammonium molybdate tetrahydrate (Fisher Scientific) and cobalt nitrate (Alfa Aesar). Crushed and sieved 100 ∼ 500 μm particles of ZrO2 were employed, which were supplied by Saint-Gobain. After incipient impregnation with aqueous solutions of the Co and Mo precursors, the supported catalysts were dried overnight at 80 °C. The dried catalysts were then calcined at 500 °C in air for 5 h.

Characterization of n ML CoMo/ZrO2 catalysts

Fig. 1 shows the TEM images of n ML CoMo-S catalysts. The inset of Fig. 1a illustrates the presence of 2.8 Å and 3.7 Å lattice fringes in the 0.5 ML catalyst, which correspond to the (-111) and (110) planes of ZrO2. The small dark dots (white arrow) suggested the presence of CoMo-S, indicating that CoMo-S species partially covered the ZrO2 support since the Co and Mo content in the 0.5 ML catalyst was insufficient to coat entire ZrO2 surface. The inset of Fig. 1b demonstrates less distinct

Conclusions

A series of CoMo catalysts supported on ZrO2 were synthesized at different CoMo-O surface coverage ranging from 0.5 to 4 ML. Highly dispersed CoMo-S species at monolayer coverage and increased ordering of CoMo-S layers at multilayer coverage were observed by TEM, XRD, and Raman spectroscopy in sulfided n ML CoMo-S catalysts. The WGS activity of these n ML catalysts increased with the surface coverage up to theoretical monolayer and then decreased for the 4 ML catalyst. The atomic S/Mo ratios

Acknowledgment

The authors gratefully acknowledge funding support from the Ohio Coal Development Office (OCDO) through the grants OCRC-R-14-11 and OCRC-R-16-11.

References (69)

  • D.S. Reichmuth et al.

    Int. J. Hydrogen Energy

    (2013)
  • M.D. Paster et al.

    Int. J. Hydrogen Energy

    (2011)
  • A. Molino et al.

    Fuel

    (2013)
  • W. Chen et al.

    Int. J. Hydrogen Energy

    (2008)
  • A.A. Ahmad et al.

    Renewable Sustainable Energy Rev.

    (2016)
  • Y. Choi et al.

    J. Power Sources

    (2003)
  • P. Hou et al.

    J. Catal.

    (1983)
  • M. Shetty et al.

    J. Catal.

    (2015)
  • L. Kaluža et al.

    Appl. Catal. A Gen.

    (2007)
  • D. Nikolova et al.

    Appl. Catal. A Gen.

    (2006)
  • M. Łaniecki et al.

    Appl. Catal. A Gen.

    (2000)
  • K.V.R. Chary et al.

    J. Catal.

    (2004)
  • X. Xie et al.

    Appl. Catal.

    (1991)
  • H. Schweiger et al.

    J. Catal.

    (2002)
  • E. Rodríguez-Castellón et al.

    Fuel

    (2008)
  • R. López Cordero et al.

    Appl. Catal. A Gen.

    (2000)
  • M. Jia et al.

    Appl. Catal. A Gen.

    (2005)
  • J. Mi et al.

    Appl. Catal. A Gen.

    (2019)
  • B. Liu et al.

    Ranliao Huaxue Xuebao J. Fuel Chem. Technol.

    (2018)
  • A.D. Gandubert et al.

    Catal. Today

    (2008)
  • R.T. Figueiredo et al.

    Catal. Today

    (2005)
  • J. Chen et al.

    Int. J. Hydrogen Energy

    (2018)
  • X. Zhang et al.

    Fuel

    (2016)
  • T. Ono et al.

    J. Catal.

    (1996)
  • B. Wang et al.

    J. Energy Chem.

    (2014)
  • B. Yoosuk et al.

    Chem. Eng. Sci.

    (2012)
  • B. Yoosuk et al.

    Catal. Today

    (2008)
  • P. Afanasiev

    Appl. Catal. A Gen.

    (2006)
  • S.S. Hla et al.

    Int. J. Hydrogen Energy

    (2011)
  • D. Ke et al.

    Int. J. Hydrogen Energy

    (2015)
  • X. Shi et al.

    Appl. Catal. A Gen.

    (2009)
  • S.L. González-Cortés et al.

    Appl. Catal. A Gen.

    (2006)
  • P. Li et al.

    Appl. Catal. A Gen.

    (2017)
  • I. Alstrup et al.

    J. Catal.

    (1982)
  • Cited by (8)

    • Hydrodeoxygenation of guaiacol via in situ H<inf>2</inf> generated through a water gas shift reaction over dispersed NiMoS catalysts from oil-soluble precursors: Tuning the selectivity towards cyclohexene

      2022, Applied Catalysis B: Environmental
      Citation Excerpt :

      Many studies exhibited higher efficiency of hydrogenation with thus generated in situ hydrogen as compared to that for ex situ one [9,20–22]. To enhance the stability of supported sulfide catalysts to water vapor, the TiO2-Al2O3, ZrO2, TiO2, Y-zeolite supports were used instead of alumina [13,14,17]. The activity and stability were found to be strongly depend on CO flow rate and water content.

    • Pd-promoting reduction of zinc salt to PdZn alloy catalyst for the hydrogenation of nitrothioanisole

      2021, Journal of Colloid and Interface Science
      Citation Excerpt :

      However, the phosphorus sources will be decomposed to highly toxic phosphine gas in the preparation of these metal phosphides [17,21,22]. For the sake of alleviating this phenomenon, the second or third metals were added into noble metal for synthesis of metal alloy catalyst [23–27]. Lakhapatri and Abraham [28] have reported that addition of Rh to Ni-based catalyst benefits in lowering the amount of sulfur adsorbed on the catalyst surface, so that the sulfur-tolerance of the catalyst is improved.

    • Effects of niobium addition on active metal and support in Co–CeO<inf>2</inf> catalyst for the high temperature water gas shift reaction

      2021, Journal of Industrial and Engineering Chemistry
      Citation Excerpt :

      Hydrogen, as a renewable form of energy, can effectively solve the problems related to energy shortage and the environmental pollution associated with the use of other energy forms [1–7]. Generally, hydrogen can be produced by reforming fossil fuels, such as, natural gas, crude oil, and coal [8–12]. However, owing to the depletion of these non-renewable energy sources, there is an urgent need for techniques that allow the production of hydrogen from renewable resources, i.e., alternative resources for the production of hydrogen are required [13–17].

    View all citing articles on Scopus
    View full text