Elsevier

Applied Catalysis A: General

Volume 569, 5 January 2019, Pages 20-27
Applied Catalysis A: General

Catalytic transformation of ethylene to propylene and butene over an acidic Ca-incorporated composite nanocatalyst

https://doi.org/10.1016/j.apcata.2018.10.017Get rights and content

Highlights

  • A composite nanocatalyst surpassed ZSM-5 and SAPO-34 counterparts in the direct ethylene to propylene (ETP) reaction.

  • The maximum ethylene conversion of 77.8% and the largest propylene yield of 34.6 wt% were obtained over the new catalyst.

  • Highest selectivity of 78.7 wt% toward propylene and butenes was obtained at 823 K with the new nanocatalyst.

  • The new catalyst possessed larger micropores and mesopores and a higher mesoporosity compared to the other two catalysts.

  • A good compromise between the amount and strength of acid sites provided the improved active centers for the ETP reaction.

Abstract

The direct conversion of ethylene to propylene (ETP reaction) has become critically important due to the growing demands for propylene as one of the key building blocks in the petrochemical industries and the depletion of appropriate resources for propylene production. In the present study, an acidic metal-incorporated composite nanocatalyst has been prepared and characterized using XRD, FTIR, FESEM, EDX, NH3-TPD, and N2 physisorption analyses and applied to the ETP reaction. The maximum conversion of 77.8% was achieved over the catalyst at 723 K. The highest propylene yield (34.6 wt%) was obtained at 773 K, however. An increasing trend with temperature (573–823 K) was evident for the total light olefins selectivity, which amounted to 78.7 wt% at 823 K. The reaction rate increased constantly to 258 mol/(kg s) with the increase in the ethylene partial pressure. The remarkable performance of the composite catalyst in terms of propylene and butenes yields and its distinguished stability during the ETP reaction compared to the closely relevant HZSM-5 and HSAPO-34 counterparts, was attributed to both an appropriate porosity and a good compromise between the strength and amount of the acidic sites to an interestingly one-mode moderate acidity.

Introduction

Light olefins (ethylene, propylene, and butene) are the key building blocks in the petrochemical industries [1]. These platform chemicals are produced industrially mainly through steam cracking [[1], [2], [3], [4], [5], [6]], catalytic cracking [[7], [8], [9]], and dehydrogenation [10,11]. To these pathways should be added the selective dimerization of ethylene to 1-butene [[12], [13], [14], [15], [16], [17], [18], [19], [20], [21]]. Hence, a lot of research is still being generated on the related catalysts [10,[22], [23], [24], [25], [26]].

The growing demand for propylene (the average of 5.7% per annum) along with the reduction in the availability of conventional sources for three to four carbon olefins have lead to the emergence of relevant on-purpose technologies to fulfill the need for the ample derivatives of propylene [[3], [4], [5], [6],22,[27], [28], [29]]. The frequent discoveries of rich gas recourses and the new technologies presented for natural gas processing in areas such as the Middle East, the Mid-Atlantic States, Russia, and the South West of the USA have led to surplus in light alkanes such as ethane which is regarded as a low-cost and available feedstock for the steam cracker units. This potential provides the largest cost advantage for the petrochemical industry [30]. Perhaps, the sole disadvantage of ethane as a feedstock is its limited product distribution (ethylene and some fuel) [30,31]. Having all of these in mind, an integrated scheme of ethane to ethylene followed by ethylene to propylene could be quite attractive by coining a new pathway that applies ethane as its starting material and ends at more demanded products such as propylene [22,32]. In this sense, the direct conversion of ethylene to propylene (ETP) or a combination of valuable chemicals including butenes and higher oligomers is emerged as one of the on-purpose technologies with the capability of responding to the displacement of the balance between supply and demand of light olefins in a flexible manner. A comprehensive overview of heterogeneous catalysts for the relevant gas-phase conversion pathways was presented in two recent review papers [22,33] that addressed all aspects of the relevant systems including the catalytic performances, productivities, life times, chemistry of active sites, etc. At variance with the relatively well-established technologies that include butene as a co-feedstock, the direct process with ethylene as the sole feedstock normally produces butene as a valuable by-product. Notably, the ETP process is viable through different reaction scenarios, such as, metathesis and oligomerization [22,34]. The former is carried out over transition metal catalysts such as Re, Mo, W, Ni and Ru [28,[34], [35], [36], [37]]. However, the process suffers from several drawbacks including the reaction intricacies and relatively quick catalyst poisoning due to the sensitivity to different types of impurities such as sulfur, water, and other sorts of hydrocarbons [22]. Taking into account these complexities, the latter pathway becomes more preferred, with the control of the reactions being relatively facile. Either way, the ETP process is known to be highly favorable thermodynamically in such a manner that, below 900 K, the standard Gibbs free energy of the oligomerization-cracking reactions is negative and below 600 K, the reaction is at complete equilibrium [38,39]. The oligomerization pathway is efficiently performed over acidic micro- and meso-porous materials such as HZSM-5, HMOR, HY, HSAPO-34, HMCM-41 and so forth [28,40]. Among those, HZSM-5 and HSAPO-34 are the most frequently applied catalysts for producing light olefins due to their high selectivity toward light olefins, fine pores which prohibit the formation of higher olefins, long lifetime particularly after metal-modification, thermal stability, shape selectivity, and also adjustable acidity [10,[41], [42], [43]]. By far, several catalysts have been tested for this purpose and improvements have been made in the relevant catalyst systems. Examples include the incorporation of various metals such as Cr, Se, and Ni into the catalyst structure, application of composite catalysts such as ZSM-5/SAPO-34 with improved characteristics to the MTO reaction [44], and the composite catalyst of HAlZSM-5 added to silica-alumina matrix for ethene and propene oligomerization [45]. Nevertheless, more investigations are required to develop highly efficient catalysts for this process in terms of propylene selectivity, productivity, and catalyst regenerability at improved conditions, e.g., the appropriate dispersion and strength of the acidic sites on the surface of the catalyst [8,11,28,36,39,46].

Having these in mind and to further advance the field, the focus of the present study was the direct conversion of ethylene to propylene (and butene) through an oligomerization-cracking pathway over a recently developed proprietary composite catalyst called GNM-1, which is characterized by the means of XRD, FTIR, FESEM, EDS, N2 physisorption, and NH3-TPD techniques. Both the reaction temperature and partial pressure of ethylene were altered to probe the catalytic performance of GNM-1 in the ETP reaction.

Section snippets

Catalyst preparation

The HZSM-5 catalyst was prepared following the well-established recipes [47,48]. A high-silica ZSM-5 sample (close to silicalite-1) was also prepared according to the same procedure except that the aluminum source was almost negligible. Typically, a clear solution of sodium aluminate and the same amount of sodium hydroxide in water was added to tetrapropylammonium hydroxide (TPAOH, 40 wt%). Then, tetraethylorthosilicate (TEOS) was slowly added and the mixture was left with stirring for 7200 s

X-ray powder diffraction

Fig. 2 displays the X-ray diffraction patterns of the three catalysts. The diffractograms of HZSM-5 and HSAPO-34 were typical patterns expected from the reference books [52,53]. For the GNM-1 sample, the multiple indicator peaks in the XRD diffractogram were attributable to a combination of different phases including pentasil (MEL, MFI) aluminum silicates, calcite, aluminum phosphate, and Ca-CHA zeolite structure with a descending order of quantity, which clearly represented the composite

Conclusion

The activity of the new acidic composite nanomaterial (GNM-1) was compared against reference HZSM-5 and HSAPO-34 samples for the ETP reaction. The XRD patterns revealed that the new composite catalyst comprised of pentasil (MEL, MFI) aluminum silicates, calcite, Ca-CHA zeolite, and aluminum phosphate, with a descending order of quantity. The NH3-TPD analyses uncovered the presence of acidic sites of medium strength on GNM-1 compared to those of the other closely relevant benchmark catalysts

Acknowledgments

The authors gratefully acknowledge the assistance from Mr. Vahid Farzaneh and the support received from IPPI under grant 53791101.

References (81)

  • R. Karimzadeh et al.

    Chem. Eng. Res. Des.

    (2009)
  • M. Ghashghaee et al.

    Chem. Eng. Res. Des.

    (2011)
  • M. Jafari Fesharaki et al.

    J. Anal. Appl. Pyrol.

    (2013)
  • M. Ghashghaee et al.

    Microporous Mesoporous Mater.

    (2013)
  • R.D. Andrei et al.

    J. Catal.

    (2015)
  • B. Nkosi et al.

    Appl. Catal. A Gen.

    (1997)
  • J.R. Sohn et al.

    Appl. Catal. A Gen.

    (2001)
  • F.T.T. Ng et al.

    Appl. Catal. A Gen.

    (1994)
  • J. Ye et al.

    J. Catal.

    (2017)
  • A. Martínez et al.

    Appl. Catal. A Gen.

    (2013)
  • L. Alvarado Perea et al.

    J. Catal.

    (2013)
  • E. Epelde et al.

    Appl. Catal. A Gen.

    (2014)
  • M. Ghashghaee

    J. Anal. Appl. Pyrol.

    (2015)
  • E. Mazoyer et al.

    J. Catal.

    (2013)
  • M. Ghashghaee et al.

    Comp. Mater. Sci.

    (2018)
  • E. Mazoyer et al.

    J. Mol. Catal. A Chem.

    (2014)
  • A.N. Mlinar et al.

    J. Catal.

    (2012)
  • H. Oikawa et al.

    Appl. Catal. A Gen.

    (2006)
  • J. Li et al.

    Propylene production by co-reaction of ethylene and chloromethane over SAPO-34

  • H.-J. Chae et al.

    J. Phys. Chem. Solids

    (2010)
  • P. Yarlagadda et al.

    Appl. Catal.

    (1990)
  • A.N. Mlinar et al.

    J. Catal.

    (2012)
  • M.Y. Kustova et al.

    Appl. Catal. B Environ.

    (2006)
  • K.Y. Lee et al.

    Appl. Catal. A Gen.

    (2009)
  • M. Ghashghaee et al.

    Appl. Catal. A Gen.

    (2017)
  • L. Regli et al.

    Stud. Surf. Sci. Catal.

    (2005)
  • A. Izadbakhsh et al.

    Microporous Mesoporous Mater.

    (2009)
  • W.A. Khanday et al.

    Arab. J. Chem.

    (2014)
  • J. Szanyi et al.

    Microporous Mesoporous Mater.

    (1996)
  • F. Lónyi et al.

    Microporous Mesoporous Mater.

    (2001)
  • G. Jiang et al.

    Appl. Catal. A Gen.

    (2008)
  • M. Ghashghaee et al.

    Microporous Mesoporous Mater.

    (2011)
  • J.C. Groen et al.

    Microporous Mesoporous Mater.

    (2003)
  • C.G.V. Burgess et al.

    J. Colloid Interface Sci.

    (1970)
  • E. Epelde et al.

    Appl. Catal. A Gen.

    (2017)
  • E. Epelde et al.

    Microporous Mesoporous Mater.

    (2014)
  • H. Zhou et al.

    Appl. Catal. A Gen.

    (2008)
  • F. Liu et al.

    Microporous Mesoporous Mater.

    (2017)
  • T. Bai et al.

    J. Energy Chem.

    (2016)
  • M. Ghashghaee et al.

    Ind. Eng. Chem. Res.

    (2018)
  • Cited by (13)

    • Promoting di-isobutene selectivity over ZnO/ZrO<inf>2</inf>-SO<inf>4</inf> in isobutene oligomerization

      2021, Chinese Journal of Chemical Engineering
      Citation Excerpt :

      Recently, Wang et al. outlined that the substitution of the Brönsted acid sites for Lewis acid sites by elements exchange significantly enhanced the ability for butene oligomerization degree control [24]. The control of Lewis acid and Brönsted acid concentrations over Ca-incorporated zeolite exhibited better catalytic transformation of C4 olefins [25]. We conjectured that combining the advantages of both Brönsted acid sites and Lewis acid sites was supposed to be a proper method to simultaneously realize the high selectivity of di-isobutene and high conversion of isobutene.

    • (Bio)Propylene production processes: A critical review

      2021, Journal of Environmental Chemical Engineering
      Citation Excerpt :

      On the In2O3-based catalysts, the rate-limiting step of overall reaction from ethanol to propylene is the acetone-to-propylene process, while acetone formation from acetaldehyde is more favorable through the aldol reaction pathway via direct coupling of two acetaldehyde than the acetate-ketonization pathway [115]. In the case of acid catalysts, it seems clear that ethylene is a main intermediate in propylene formation [116–120]. Several authors showed that diethyl ether is the first product of ethanol dehydration at low temperature, which can crack to form ethylene [68,121].

    • Propene production at low temperature by bimetallic Ni-Mo and Ni-Re catalysts on mesoporous MCM-41 prepared using template ion exchange

      2021, Fuel
      Citation Excerpt :

      Therefore, there is uncertainty how the chemical process industry will reduce the gap between offer and demand of this valuable chemical. One of the most interesting options to produce propene is the direct conversion of ethene to propene [3–20]. This process involves the production of propene by feeding only ethene diluted in nitrogen as reactant without mixing it with other type of hydrocarbons.

    • Dual role of ferric chloride in modification of USY catalyst for enhanced olefin production from refinery fuel oil

      2019, Applied Catalysis A: General
      Citation Excerpt :

      Light olefins (C2H4, C3H6, and C4H8) are the largest volume building blocks for many chemicals and petrochemicals [1–3]. These starting materials are produced industrially via steam cracking [4–9], catalytic cracking [10–13], metathesis [14], dehydrogenation [15–17], and dimerization [18–20]. Over the years, intensive research has been conducted on converting heavy and extra-heavy hydrocarbons to products of growing worldwide demands, such as light olefins and liquid fuels [5,9,11,21,22].

    View all citing articles on Scopus
    View full text