Next Article in Journal / Special Issue
Interlayer Properties of In-Situ Oxidized Porous Stainless Steel for Preparation of Composite Pd Membranes
Previous Article in Journal
Effect of the Chemical Composition of Mesoporous Cerium-Zirconium Oxides on the Modification with Sulfur and Gold Species and Their Application in Glycerol Oxidation
Previous Article in Special Issue
Hydrogen and Deuterium Solubility in Commercial Pd–Ag Alloys for Hydrogen Purification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrogen Utilization in Green Fuel Synthesis via CO2 Conversion to Methanol over New Cu-Based Catalysts

by
Lorenzo Spadaro
1,2,*,
Mariarita Santoro
2,
Alessandra Palella
1 and
Francesco Arena
1,2
1
Institute CNR-TAE “Nicola Giordano”, Via Salita S. Lucia 5, I-98126 Messina, Italy
2
Dipartimento di Ingegneria, Università di Messina, Viale F. Stagno D’Alcontres 31, I-98166 Messina, Italy
*
Author to whom correspondence should be addressed.
ChemEngineering 2017, 1(2), 19; https://doi.org/10.3390/chemengineering1020019
Submission received: 10 November 2017 / Revised: 6 December 2017 / Accepted: 14 December 2017 / Published: 19 December 2017

Abstract

:
The use of hydrogen as an energy vector and raw material for “very clean liquid fuels” manufacturing has been assessed by the catalytic conversion of CO2 to methanol over copper based catalysts. A systematic evaluation of copper based catalysts, prepared varying the chemical composition, has been carried out at 0.1–5.0 MPa of total pressure and in the range of 453–513 K by using a semi-automated LAB-microplant, under CO2/H2 reactant mixture (1/3), fed at GHSV of 8.8 NL∙kgcat−1∙h−1. Material’s properties have been investigated by the means of chemical-physical studies. The findings disclose that the addition of structure promoters (i.e., ZrO2/CeO2) strongly improves the textural properties of catalysts, in term of total surface area and exposure of metal surface area (MSA), also reducing the sintering phenomena. The results of the catalytic study clearly prove a structure-activity relationship at low reaction pressure (0.1 MPa), while at higher pressure (3.0–5.0 MPa) the reaction path is insensitive to structure and chemical composition.

1. Introduction

The combined use of CO2 and renewable H2, provided by electrolysis or water splitting, in the chemicals and fuels manufacturing, nowadays looks very attractive. In particular, CO2 conversion into fuel products is the most promising way for a larger H2 utilization, as an energy vector, providing for instance methanol and other light oxygenated hydrocarbons for fuel cell (FC) applications (i.e., direct methanol/alcohols FC). Indeed, the catalytic hydrogenation of CO2 is still more effective than direct photo-conversion [1,2,3,4,5,6,7,8,9].
On the other side, the production of methanol via the hydrogenation of CO2, using electro-catalytically generated H2, is of particular interest also for renewable energy concepts, since it can serve as a way to convert and store the excess electrical energy into chemical energy, contributing to smooth the natural fluctuation in the supply of renewable energy [10,11]. Moreover, the CO2 hydrogenation can contribute to curb CO2 emissions, which are a cause of the increase in global temperatures and climate change due to the “greenhouse effect” [12].
Prompted by seeking better performing formulations, many authors have explored new catalytic compositions, active in the CO2 hydrogenation, alternative to the copper-based, such as combinations of Pd [13,14], Au [10,15], or Mo2C [16,17]. Nevertheless, copper-based catalysts still remain the most suitable catalysts for commercial applications [18]. Cu-based catalysts suffer from detrimental effects of the water produced during catalyst activation and reaction, which inhibits catalytically active sites through the irreversible oxidation of Cu+ species to Cu2+, leading to a loss of interfacial area and accelerating the catalyst sintering and the synthesis rate decay [18]. On this address, in the past decades, significant effort has been devoted to improve the properties of Cu-based catalysts by the use of promoters and carriers/supports [19,20,21,22,23]. Several metal oxides are reported in literature for improving the catalytic performance of copper-based catalysts, in term of activity, selectivity, and stability [24,25,26,27,28,29], also enhancing CO2 adsorption with respect to un-promoted systems. In particular, in our previous works [30,31,32,33,34], we disclose the positive effects of ZrO2 on the catalytic performance of Cu-ZnO catalyst with respect to that of Al2O3-supported, as a result of the lower hydrophilic nature of ZrO2 than Al2O3, which facilitates water desorption, and its property of stabilizing Cu+–O species, which favor CO2 adsorption/activation and methanol selectivity. Then, in more recent studies, we have clearly assessed that ZrO2 affects structure/morphology and texture of Cu-ZnO based catalysts, while cerium oxides can divalently act as an electronic promoter and improver of surface functionality of Cu phase [33,34].
Therefore, the present work aims to study the combined effects of ZrO2 and CeO2 addition on the chemical-physical properties of copper based catalyst, in efforts to shed light on the CO2 hydrogenation reaction pathways.

2. Materials and Methods

2.1. Synthesis of the Catalysts

Cu–ZnO based catalysts, with a Cu/Zn atomic ratio of circa 3, were prepared by reverse coprecipitation under ultrasound irradiation route, varying ZrxCe(1−x)O2 carrier composition between 0.2 < x < 0.95, according to procedure elsewhere described [30]. Table 1 shows the list of catalysts, also reporting the chemical-physical properties.

2.2. Characterization of the Catalysts

Surface area (SA), pore volume (PV), and average pore diameter (APD) were obtained from nitrogen adsorption/desorption isotherms at 77 K, using ASAP 2020, Micromeritics Instrument. Isotherms were elaborated by the standard BET and BJH methods for SA and PV evaluation, respectively.
X-ray diffraction (XRD) analysis of samples in the 2θ range 10–80° was performed using a Philips X-Pert diffractometer operating with Ni β-filtered Cu Kα radiation at 40 kV and 30 mA.
Metal Surface Area (MSA) values were obtained by measurements of N2O “single-pulse” titrations, assuming a Cu:N2O = 2 titration stoichiometry [35,36]. Before measurements, catalysts were reduced in situ at 573 K in H2 flow (100 STP mL/min) for 1 h. After reduction, the samples were “washed” in He carrier flow at 583 K for 15 min, and then cooled to 363 K.

2.3. Catalytic Measurements

Catalytic tests were performed at 0.1–5.0 MPa and 453–513 K in a semiautomatic Laboratory microplant, equipped with a AISI 316 stain steel Plug Flow Reactor (i.d., 10 mm; e.d., 12 mm; l., 250 mm), loaded with 1.0 g of catalyst diluted (1/1, wt/wt) with quartz granules (2 mm). Before catalytic tests, catalysts were in-situ activated at 573 K in reducing atmosphere, flowing (5.5 NL·g−1·h−1) 5% H2/Ar gas mixture for 1 h at a heating rate of 10 K·min−1. After catalysts activation the reactor was fed with a CO2/H2/N2 (3/9/1) mixture flowing at 80 STPcm3·min−1 (GHSV, 8.800 NL∙kgcat−1∙h−1) using nitrogen as internal standard for GC analysis. The composition of inlet and outlet streams was analyzed in real time by an inline gas chromatograph HP 6890 equipped with a TCD detector and two packed columns (Porapack Q and 5A molecular sieve, Sigma-Aldrich (Milano, Italy)) connected in series for the analysis of permanent gasses, and a FID detector coupled with a capillary column (PoraPLOT Q, i.d., 0.53 mm; l, 30 m, Agilent (Cernusco sul Naviglio (MI), Italy)) for the hydrocarbon species quantification. Methanol and CO were almost the only products formed, while methane and DME are detected only at trace levels (<0.01 mol/mol %).

3. Results

3.1. Chemical-Physical Characterization

As shown in Table 1 and Figure 1, the progressive replacement of Ce3+/4+ ions with Zr4+ (i.e., ZrxCe(1-x)O2; 0.2 < x < 0.95) strongly affects both structural and morphological properties of catalysts. Indeed, a remarkable rise of both total and metallic surface area was reported with a consequent increase of the copper–zinc oxide interfacial area, accompanied by an also improved resistance to sintering due to the activation procedure (reduction).
As reported in our previous work [30,31,32], catalyst activation consists in a controlled reduction to obtain active metallic copper. The action of temperature and H2 for oxygen removal (MeOx + xH2 → xH2O; Me = Cu and Ce) leads to the reconstruction of catalyst which can also drive to a higher crystallization with a consequent loss of total surface area (sintering). On this address, the effects of the activation procedure were fully investigated in our previous works [30,31,32].
In particular, the catalyst without ZrO2 (CuZn_CZ_0) displays the lowest value of surface area (47 m2·g−1), which further decreases by 28% after reduction (34 m2·g−1). By contrast, the system with no CeO2 (CuZn_CZ_100) shows the highest exposure of total surface (~154 m2·g−1) and the lowest decay for the reduction treatment (ΔSA = 14%), confirming the strong effect of zirconia as structural promoter (Figure 1). Indeed, along the replacement of Ce ions in the lattice, the total surface area grows from 67 m2·g−1 to 125 m2·g−1, also enhancing MSA value (14–28 m2·g−1), which mirrors an improved dispersion of active phase, confirmed by an increase of the MSA/SAR ratio. Then, Figure 2 displays the diffraction patterns of catalysts as prepared (a) and after reduction treatment (b). Irrespective of chemical formulation, XRD analysis signals the lack of “long-range” crystalline order in all samples, due to the ultrasonic preparation method, Figure 2a. In particular, the XRD profile is characterized for all catalysts by the presence of two broad peaks, placed at ca. 31 and 56 2θ degrees, with an increasing intensity at the raise of Zr content. Therefore, catalyst reduction leads to a bulk reconstruction, as shown in Figure 2b. Namely, the catalysts at higher Zr content exhibit the signals of copper in the metallic state, whose most intense reflections appear at 43.3° (<111>) and 50.4° (<200>), respectively. In addition, the occurrence into CuZn_CZ_90 catalyst of sharp peaks, at 35.68° and 38.88°, diagnostics the presence of isolate Tenorite phase (CuO) [JCPDS 5–661], while in the XRD pattern of the reduced CuZn_CZ_95 sample are also evident the diffraction signals of cubic Zirconia [JCPDS 27-0997]. On the other hand, the reflection peaks of metallic Cu are less marked in the reduced samples at lower Zr content (i.e., CuZn_CZ_50 and CuZn_CZ_20), probing a stronger re-oxidation, due to the greater oxygen storage capacity of CeO2. Furthermore, small intense lines of Zincite phase (ZnO) are detected in the XRD profile of reduced CuZn_CZ_50 catalyst. Therefore, the good chemical stability of catalysts was further proved by results of catalytic test and by XRD analysis on samples after catalytic tests (data not reported here for the sake of brevity).

3.2. Catalytic Activity

The results of catalytic tests in the CO2 hydrogenation in the range of 453–513 K and at 0.1, 3.0 and 5.0 MPa are summarized in Table 2 in terms of H2 conversion ( X H 2 %), selectivity to methanol ( S CH 3 OH %) and reaction rate (mmolH2·h−1·gcat−1).
As extensively reported in literature, CO2 hydrogenation drives to methanol and CO over Cu-based catalyst through the reactions of methanol synthesis (Equations (1) and (2)) and Reverse Water Gas Shift (Equation (3)), while the methane formation (Equation (4)) is strongly un-favored under such reaction conditions [27,37]:
CO2 + 3H2 ⇆ CH3OH + H2O
CO + 2H2 ⇆ CH3OH
CO2 + H2 ⇆ CO + H2O
CO2 + 4H2 ⇆ CH4 + 2H2O
In respect of this, a progressive increase of H2 conversion, associated with a significant decrease in methanol selectivity, has been observed with the increase in temperature, independent of chemical composition and pressure. According to the greater partial pressure of reactants on the catalyst surface, the H2 conversion increases further over all samples by upping the pressure from 0.1 MPa to 5.0 MPa. As a rule, ZrO2 content growth leads to a higher hydrogenation of CO2, although, the selectivity to methanol is notably reduced.
In particular, reaction rate raises from 8.7 mmol·h−1·gcat−1 to the greater 20.1 mmol·h−1·gcat−1 at the highest temperature (513 K), moving from 0.1 MPa to 5.0 MPa by using the most active catalyst (CuZn_CZ_100). While the reaction rate increases only in the range of 2.3–7.8 (mmol·h−1·gcat−1), at the same test condition, employing the less active catalyst (CuZn_CZ_0).
In catalysis, “Activation Energy” counts for the energy that must be supplied to catalyst to observe any chemical reaction. On this account, a further evaluation of the effects of carrier composition on the catalytic functionality of Cu-ZnO systems can be done comparing the values of Apparent Activation Energy (Eapp (kJ·mol−1)), obtained from the Arrhenius’s plot of differential conversion data at 0.1 MPa ( X H 2 < 10%, T = 453–513 K), reported in Figure 3 [30,31,38]. Indeed, the progressive increase of CeO2 content leads to a decrease in Eapp value from 97 kJ·mol−1 of CuZn_CZ_0 sample to 26 kJ·mol−1 of CuZn_CZ_100 system, as proof of the promoter effect of ceria. The catalytic data at higher pressure (i.e., 3.0 and 5.0 MPa) were not considered for the evaluation of Eapp, because the hydrogen conversion ( X H 2 ) overcame the value of 10%.
As well known, at the kinetic regime, the value of the Apparent Activation Energy (Eapp) is directly related to the kinetic constant “K” and, consequently, to the reaction rate (rate ∝ K) through Arrhenius’s Law, also reflecting the energies needed for the chemisorption and desorption processes of the molecule on the surface of active sites [39,40]. Therefore, the decreasing in the value of Eapp observed with the raise of Ceria contents in the catalyst formulation can also mirror a more favorable chemisorption or desorption steps in the hydrogenation reaction.

4. Discussion

CO2 hydrogenation to methanol by using Cu-based catalytic systems promoted by several oxides/carriers has been widely investigated in literature [3,5,16,17,18,19,20,21,22,23,24,27,28,29,30,31,32], although very few examples of multicomponent CuO-ZnO-CeO2 and CuO-ZnO-CeO2-ZrO2 systems are still present [25,26,33,34,41,42].
Cerium oxides are well recognized catalytic materials for the oxygen storage capacity of Ceria and for several features such as to be an enhancer of active phase dispersion, for weakening the C–O bond and for improving the chemisorption of H2 [43,44,45]. As an electronic promoter, Ceria also plays a crucial role in the behavior of the catalyst, thanks to its peculiar reactivity [43,44]. In this work, the effect of catalyst formulation becomes more clear by plotting the logarithm of H2 conversion (lg X H 2 ) against the selectivity to methanol, Figure 4. Namely, at 0.1 MPa the catalytic behavior of all catalysts can be described through three different curves of activity, which respond to different catalyst formulation. Differently, at the higher pressure (3.0 and 5.0 MPa) the samples exhibit similar catalytic paths. These findings suggest that catalyst chemisorption capacity, also with respect to intermediate products, could affect the catalytic behavior at low reactants coverage. In particular, at 0.1 MPa the catalyst with composition “Cu-ZnO-ZrO2” (CuZn_CZ_100) promotes mainly the reaction of hydrogenation which leads to form CO. By contrast, the catalyst with the chemical formulation “Cu-ZnO-CeO2” (CuZn_CZ_0) favors mainly the methanol synthesis, perhaps, thanks to the oxygen storage capacity of Ceria. Indeed, according to the reaction mechanism in which methanol synthesis occurs via hydrogenation of a “bridge-formate” intermediate [46,47], CuZn_CZ_0 would promote methanol because of its higher capacity to weaken and bend C–O bonds. In addition, the higher hydrogen coverage on the surface, due to the H2-spillover phenomena, could further enhance the selectivity to methanol for CuZn_CZ_0 catalyst. Therefore, the catalytic behavior held by the catalysts having both ZrO2 and CeO2 promoter/carrier results in a right equilibrium of chemical-physical properties of the two different oxides, being a good compromise between the two extreme catalytic performance.
Therefore, the effects of ceria, as structural and electronic promoter, can be both at level of solid-state interactions and reactivity toward reactant gas [30,31,32,33,34,45,48], affecting the redox and electronic properties of the active phase and modelling the adsorption behavior. On the other hand, the enhanced reducibility (data not here reported) and the well stated surface affinity for H2 can further justify the lower value of Apparent Activation Energy [30,31,32,33,34,45,48].
Then, to shed light on structure-activity relationship, the catalytic results were analyzed according to their textural properties (SA and MSA). On this account, Figure 5a,b reports reaction rate and “normalized methanol yield” at 5.0 MPa and at varying of MSA and MSA/SAR values, respectively. In particular, the CO2 hydrogenation rate increases almost proportionally with the increase of metal exposure, Figure 5a, according to the consistent rise and reconstruction of the active phases/sites (Cu0 ⇆ Cu+). On the other side, Figure 1 and Table 1 shown that the value of MSA increases by the growth of the atomic fraction of Zr, as proof of the greater efficacy of ZrO2 carrier as structural promoter and improver of metal dispersion. Moreover, methanol synthesis over Cu-based catalyst is characterized by a “dual site mechanism”, which invokes a right equilibrium between Cu+ and Cu0 active sites, according to the following equation 5:
ratemethanol = K · [CO2] · [H2] ∝ K· [Cu(I)] · [Cu(0)]
where a greater concentration of Cu0 sites drive to high hydrogen supply on the catalytic surface, while a major exposition of Cu+ active sites on SA strongly enhance CO2 chemisorption. Therefore, the catalyst performance depends on the right “deal” between the two catalytic functionalities [3,38,45,46,49,50,51].
In other words, all “Cu-ZnO-ZrO2-CeO2” catalysts obey the occurrence of a similar “dual-site” reaction path that is a mechanism with two active centers involved in the catalytic process of CO2 hydrogenation. The first one should be the oxide phase, with basic properties, which attends to adsorb and activate CO2 molecules for running hydrogenation to methanol, through the formation of a formate as intermediate [46]. In more detail, the CO2 chemisorption can occur both via Zirconia Lewis basic sites [47,49,50], as well as by action of metal alkaline ZnO phase which enhances the affinity of Cu(+)–O to CO2 [18]. As proof, in our previous works we have attested a larger CO2 uptake, consequently to a greater concentration of basic sites [33,34]. On the other hand, the other sites should accomplish the adsorption and dissociation of H2 molecules for the reaction with CO2 [24,47]. On this account, the occurring of a reaction pathway which involves the existence of two kind of reaction sites is confirmed for “Cu-ZnO-ZrO2-CeO2” catalytic systems by a “volcano-shape” relationship, obtained by plotting normalized methanol yield versus the ratio MSA/SAR, here used as an indirect index of the dual-sites equilibrium (Figure 5b). Namely, the wide range of values assumed by the normalized yield at the varying of MSA/SAR discloses the structure-activity relationship of these catalytic materials, and at the same time, reflects the effect of chemical composition. In addition, the existence of a maximum value of methanol yield, which should be achieved at 5.0 MPa at an “ideal” ratio of MSA/SAR near to 0.3, clearly signs the need of a perfect synergism between the two catalytic functionalities, available through a proper design of the catalyst [51,52,53,54].
In particular, despite the lower activity, the use of Ceria in combination with Zirconia leads to a more favorable Activation Energy (ranging from 97 to 27 kJ·mol−1), which also results in higher values of methanol yield than that of the sample with no Ceria. In other words, even the addition of small quantities of CeO2 as promoter (i.e., 5–10 wt%) confers to catalyst a peculiar reactivity towards the methanol formation, by enhancing the surface functionality of Cu-ZnO-ZrO2 system.
These findings concur well with our earlier results, especially in term of methanol yield and Eapp [30,31,32,33,34]. In addition, the catalytic performance of CuO-ZnO-CeO2-ZrO2 systems represent an advancement with respect to the results recently reported in literature by Cu-based catalysts [26,41,42].

5. Conclusions

The catalytic performance in CO2 hydrogenation process for a series of new “Cu-ZnO-ZrO2-CeO2” catalysts have been assessed, approaching real industrial process conditions.
The work assessed the effect of chemical composition and preparation method regarding the sintering phenomena due to the activation procedure (reduction) in which the action of temperature and H2 in the oxygen removing and metallic copper formation (MeOx + xH2 = xH2O; Me = Cu and Ce). This leads to catalyst reconstruction which can also drive to higher crystallization with a consequent loss of total surface area. The progressive replacement of Ce3+/4+ ions with Zr4+ (i.e., ZrxCe(1−x)O2; 0.2 < x < 0.95) has been proved affecting both SA and MSA exposure, improving also the resistance to sintering phenomena.
The progressive increase of CeO2 content has been ascertained to lead to a more favorable Activation Energy values. The results obtained with CuO-ZnO-CeO2-ZrO2 systems represent a further advancement in the comprehension of catalytic pathway of Cu-based catalyst, The “dual-site” mechanism, as reaction pathway in methanol synthesis, has been confirmed, through a “volcano-shape” relationship, obtained between hydrogenation activity and catalytic structure of “Cu-ZnO-ZrO2-CeO2” systems.
The addition of even low quantities of CeO2 (i.e., 5–10%) allows one to obtain materials with enhanced chemical properties, maintaining the striking structural promoter effect of ZrO2, and, at the same time, introducing the functionality of CeO2, with a satisfactory compromise between activity and selectivity.

Author Contributions

Lorenzo Spadaro and Francesco Arena conceived and designed the experiments and supervised the research activity. Mariarita Santoro and Alessandra Palella performed the experiments and analyzed the experimental data. All the authors participated to the discussion of the results and to the drafting of the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saeidi, S.; Amin, N.A.S.; Rahimpour, M.R. Hydrogenation of CO2 to value-added products—A review and potential future developments. J. CO2 Util. 2014, 5, 66–81. [Google Scholar] [CrossRef]
  2. Larson, E.D.; Tingjin, R. Synthetic fuel production by indirect coal liquefaction. Energy Sustain. Dev. 2003, 7, 79–102. [Google Scholar] [CrossRef]
  3. Liu, X.-M.; Lu, G.Q.; Yan, Z.-F.; Beltramini, J. Recent Advances in Catalysts for Methanol Synthesis via Hydrogenation of CO and CO2. Ind. Eng. Chem. Res. 2003, 42, 6518–6530. [Google Scholar] [CrossRef]
  4. Song, C. Global challenges and strategies for control, conversion and utilization of CO2 for sustainable development involving energy, catalysis, adsorption and chemical processing. Catal. Today 2006, 115, 2–32. [Google Scholar] [CrossRef]
  5. Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703. [Google Scholar] [CrossRef] [PubMed]
  6. Fiedler, E.; Grossmann, G.; Kersebohm, D.B.; Weiss, G.; Witte, C. Methanol. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley: New York, NY, USA, 2000. [Google Scholar]
  7. Chen, C.-Y.; Yu, J.C.-C.; Nguyen, V.-H.; Wu, J.C.-S.; Wang, W.-H.; Kočí, K. Reactor Design for CO2 Photo-Hydrogenation toward Solar Fuels under Ambient Temperature and Pressure. Catalysts 2017, 7, 63. [Google Scholar] [CrossRef]
  8. Olah, G.A.; Goeppert, A.; Prakash, G.K.S. Chemical recycling of carbon dioxide to methanol and dimethyl ether: From greenhouse gas to renewable environmentally carbon neutral fuels and synthetic hydrocarbons. J. Org. Chem. 2009, 74, 487–498. [Google Scholar] [CrossRef] [PubMed]
  9. Pontzen, F.; Liebner, W.; Gronemann, V.; Rothaemel, M.; Ahlers, B. CO2-based methanol and DME—Efficient technologies for industrial scale production. Catal. Today 2011, 171, 242–250. [Google Scholar] [CrossRef]
  10. Hartadi, Y.; Widmann, D.; Behm, R.J. Methanol formation by CO2 hydrogenation on Au/ZnO catalysts—Effect of total pressure and influence of CO on the reaction characteristics. J. Catal. 2016, 333, 238–250. [Google Scholar] [CrossRef]
  11. Schlögl, R. Chemistry’s Role in Regenerative Energy. Angew. Chem.-Int. Ed. 2011, 50, 6424–6426. [Google Scholar] [CrossRef] [PubMed]
  12. Larminie, J.; Dicks, A. Fuel Cell Systems Explained, 2nd ed.; Wiley: New York, NY, USA, 2003. [Google Scholar]
  13. Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd-Cu catalysts for selective CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2015, 170–171, 173–185. [Google Scholar] [CrossRef]
  14. Rui, N.; Wang, Z.; Sun, K.; Ye, J.; Ge, Q.; Liu, C.J. CO2 hydrogenation to methanol over Pd/In2O3: Effects of Pd and oxygen vacancy. Appl. Catal. B Environ. 2017, 218, 488–497. [Google Scholar] [CrossRef]
  15. Vourros, A.; Garagounis, I.; Kyriakou, V.; Carabineiro, S.A.C.; Maldonado-Hódar, F.J.; Marnellos, G.E.; Konsolakis, M. Carbon dioxide hydrogenation over supported Au nanoparticles: Effect of the support. J. CO2 Util. 2017, 19, 247–256. [Google Scholar] [CrossRef]
  16. Ma, Y.; Guan, G.; Hao, X.; Cao, J.; Abudula, A. Molybdenum carbide as alternative catalyst for hydrogen production—A review. Renew. Sustain. Energy Rev. 2017, 75, 1101–1129. [Google Scholar] [CrossRef]
  17. Liu, X.; Song, Y.; Geng, W.; Li, H.; Xiao, L.; Wu, W. Cu-Mo2C/MCM-41: An Efficient Catalyst for the Selective Synthesis of Methanol from CO2. Catalysts 2016, 6, 75. [Google Scholar] [CrossRef]
  18. Arena, F.; Mezzatesta, G.; Spadaro, L.; Trunfio, G. Latest advances in the catalytic hydrogenation of CO2 to methanol/dimethylether. In Transformation and Utilization of Carbon Dioxide, Green Chemistry and Sustainable Technology; Bhanage, B.M., Arai, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar] [CrossRef]
  19. Jadhav, S.G.; Vaidya, P.D.; Bhanage, B.M.; Joshi, J.B. Catalytic carbon dioxide hydrogenation to methanol: A review of recent studies. Chem. Eng. Res. Des. 2014, 92, 2557–2567. [Google Scholar] [CrossRef]
  20. Deng, K.; Hu, B.; Lu, Q.; Hong, X. Cu/g-C3N4 modified ZnO/Al2O3 catalyst: Methanol yield improvement of CO2 hydrogenation. Catal. Commun. 2017, 100, 81–84. [Google Scholar] [CrossRef]
  21. Li, S.M.M.-J.; Zeng, Z.; Liao, F.; Hong, H.; Tsang, S.C.E. Enhanced CO2 hydrogenation to methanol over CuZn nanoalloy in Ga modified Cu/ZnO catalysts. J. Catal. 2016, 343, 157–163. [Google Scholar] [CrossRef]
  22. Umegaki, T.; Kojima, Y.; Omata, K. Effect of oxide coating on performance of copper-zinc oxide-based catalyst for methanol synthesis via hydrogenation of carbon dioxide. Materials 2015, 8, 7738–7744. [Google Scholar] [CrossRef] [PubMed]
  23. Da Silva, R.J.; Pimentel, A.F.; Monteiro, R.S.; Mota, C.J.A. Synthesis of methanol and dimethyl ether from the CO2 hydrogenation over Cu/ZnO supported on Al2O3 and Nb2O5. J. CO2 Util. 2016, 15, 83–88. [Google Scholar] [CrossRef]
  24. Xiao, J.; Mao, D.; Guo, X.; Yu, J. Effect of TiO2, ZrO2, and TiO2-ZrO2 on the performance of CuO-ZnO catalyst for CO2 hydrogenation to methanol. Appl. Surf. Sci. 2015, 338, 146–153. [Google Scholar] [CrossRef]
  25. Chang, K.; Wang, T.; Chen, J.C. Hydrogenation of CO2 to methanol over CuCeTiOx catalysts. Appl. Catal. A Gen. 2017, 206, 704–711. [Google Scholar] [CrossRef]
  26. Ouyang, B.; Tan, W.; Liu, B. Morphology effect of nanostructure ceria on the Cu/CeO2 catalysts for synthesis of methanol from CO2 hydrogenation. Catal. Commun. 2017, 95, 36–39. [Google Scholar] [CrossRef]
  27. Gao, P.; Li, F.; Zhan, H.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Wang, H.; Sun, Y. Influence of Zr on the performance of Cu/Zn/Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. J. Catal. 2013, 298, 51–60. [Google Scholar] [CrossRef]
  28. Gao, P.; Li, F.; Zhan, H.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Sun, Y. Fluorine-modified Cu/Zn/Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. Catal. Commun. 2014, 50, 78–82. [Google Scholar] [CrossRef]
  29. Huang, C.; Chen, S.; Fei, X.; Liu, D.; Zhang, Y. Catalytic Hydrogenation of CO2 to Methanol: Study of Synergistic Effect on Adsorption Properties of CO2 and H2 in CuO/ZnO/ZrO2 System. Catalysts 2015, 5, 1846–1861. [Google Scholar] [CrossRef]
  30. Arena, F.; Barbera, K.; Italiano, G.; Bonura, G.; Spadaro, L.; Frusteri, F. Synthesis, characterization and activity pattern of Cu-ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J. Catal. 2007, 249, 185–194. [Google Scholar] [CrossRef]
  31. Arena, F.; Italiano, G.; Barbera, K.; Bordiga, S.; Bonura, G.; Spadaro, L.; Frusteri, F. Solid-state interactions, adsorption sites and functionality of Cu-ZnO/ZrO2 catalysts in the CO2 hydrogenation to CH3OH. Appl. Catal. A Gen. 2008, 350, 16–23. [Google Scholar] [CrossRef]
  32. Arena, F.; Italiano, G.; Barbera, K.; Bonura, G.; Spadaro, L.; Frusteri, F. Basic evidences for methanol-synthesis catalyst design. Catal. Today 2009, 143, 80–85. [Google Scholar] [CrossRef]
  33. Arena, F.; Mezzatesta, G.; Zafarana, G.; Trunfio, G.; Frusteri, F.; Spadaro, L. How oxide carriers control the catalytic functionality of the Cu-ZnO system in the hydrogenation of CO2 to methanol. Catal. Today 2013, 210, 39–46. [Google Scholar] [CrossRef]
  34. Arena, F.; Mezzatesta, G.; Zafarana, G.; Trunfio, G.; Frusteri, F.; Spadaro, L. Effects of oxide carriers on surface functionality and process performance of the Cu-ZnO system in the synthesis of methanol via CO2 hydrogenation. J. Catal. 2013, 300, 141–151. [Google Scholar] [CrossRef]
  35. Agrell, J.; Birgersson, H.; Boutonnet, M.; Melián-Cabrera, I.; Navarro, R.M.; Fierro, J.L.G. Production of hydrogen from methanol over Cu/ZnO catalysts promoted by ZrO2 and Al2O3. J. Catal. 2003, 219, 389–403. [Google Scholar] [CrossRef]
  36. Evans, J.W.; Wainwright, M.S.; Bridgewater, A.J.; Young, D.J. On the determination of copper surface area by reaction with nitrous oxide. Appl. Catal. 1983, 7, 75–83. [Google Scholar] [CrossRef]
  37. Zhang, L.; Zhang, Y.; Chen, S. Effect of promoter SiO2, TiO2 or SiO2-TiO2 on the performance of CuO-ZnO-Al2O3 catalyst for methanol synthesis from CO2 hydrogenation. Appl. Catal. A Gen. 2012, 415–416, 118–123. [Google Scholar] [CrossRef]
  38. Arena, F.; Di Chio, R.; Fazio, B.; Espro, C.; Spiccia, L.; Palella, A.; Spadaro, L. Probing the functionality of nanostructured MnCeOx catalysts in the carbon monoxide oxidation. Part I. Influence of cerium addition on structure and CO oxidation activity. Appl. Catal. B Environ. 2017, 210, 14–22. [Google Scholar] [CrossRef]
  39. Chorkendorff, I.; Niemantsverdiet, J.W. Concepts of Modern Catalysis and Kinetics; Wiley-VCH Verlag GmgH & Co. KGaA: Weinheim, Germany, 2007. [Google Scholar]
  40. Bartholomew, C.H.; Ferrauto, R.J. Foundamentals of Industrial Catalytic Processes; Wiley-Interscience: Hoboken, NJ, USA, 2006. [Google Scholar]
  41. Li, L.; Mao, D.; Yu, J.; Guo, X. Highly selective hydrogenation of CO2 to methanol over CuO-ZnO-ZrO2 catalysts prepared by a surfactant-assisted co-precipitation method. J. Power Sources 2015, 279, 394–404. [Google Scholar] [CrossRef]
  42. Angelo, L.; Kobl, K.; Tejada, L.M.M.; Zimmermann, Y.; Parkhomenko, K.; Roger, A.-C. Study of CuZnMOx oxides (M = Al, Zr, Ce, CeZr) for the catalytic hydrogenation of CO2 into methanol. C. R. Chim. 2015, 18, 250–260. [Google Scholar] [CrossRef]
  43. Arena, F.; Famulari, P.; Trunfio, G.; Bonura, G.; Frusteri, F.; Spadaro, L. Probing the factors affecting structure and activity of the Au/CeO2 system in total and preferential oxidation of CO. Appl. Catal. B Environ. 2006, 66, 81–91. [Google Scholar] [CrossRef]
  44. Arena, F.; Famulari, P.; Interdonato, N.; Bonura, G.; Frusteri, F.; Spadaro, L. Physico-chemical properties and reactivity of Au/CeO2 catalysts in total and selective oxidation of CO. Catal. Today 2006, 116, 384–390. [Google Scholar] [CrossRef]
  45. Trovarelli, A. Catalysis by Ceria and Related Materials; Imperial College Press: London, UK, 2002. [Google Scholar]
  46. Kunkes, E.L.; Studt, F.; Abild-Pedersen, F.; Schlögl, R.; Behrens, M. Hydrogenation of CO2 to methanol and CO on Cu/ZnO/Al2O3: Is there a common intermediate or not? J. Catal. 2015, 328, 43–48. [Google Scholar] [CrossRef]
  47. Fisher, I.A.; Bell, A.T. In-Situ Infrared Study of Methanol Synthesis from H2/CO2 over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 1997, 172, 222–237. [Google Scholar] [CrossRef]
  48. Spadaro, L.; Arena, F.; Granados, M.L.; Ojeda, M.; Fierro, J.L.G.; Frusteri, F. Metal-support interactions and reactivity of Co/CeO2 catalysts in the Fischer-Tropsch synthesis reaction. J. Catal. 2005, 234, 451–462. [Google Scholar] [CrossRef]
  49. Fisher, I.A.; Bell, A.T. In-Situ Infrared Study of Methanol Synthesis from H2/CO over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 1998, 178, 153–173. [Google Scholar] [CrossRef]
  50. Pokrovski, K.; Jung, K.T.; Bell, A.T. Investigation of CO and CO2 Adsorption on Tetragonal and Monoclinic Zirconia. Langmuir 2001, 17, 4297–4303. [Google Scholar] [CrossRef]
  51. Tada, S.; Watanabe, F.; Kiyota, K.; Shimoda, N.; Hayashi, R.; Takahashi, M.; Nariyuki, A.; Igarashi, A.; Satokawa, S. Ag addition to CuO-ZrO2 catalysts promotes methanol synthesis via CO2 hydrogenation. J. Catal. 2017, 351, 107–118. [Google Scholar] [CrossRef]
  52. Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.; Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2 hydrogenation to methanol. J. Catal. 2016, 343, 133–146. [Google Scholar] [CrossRef]
  53. Yang, C.; Ma, Z.; Zhao, N.; Wei, W.; Hu, T.; Sun, Y. Methanol synthesis from CO2-rich syngas over a ZrO2 doped CuZnO catalyst. Catal. Today 2006, 115, 222–227. [Google Scholar] [CrossRef]
  54. Ma, Y.; Sun, Q.; Wu, D.; Fan, W.-H.; Zhang, Y.-L.; Deng, J.-F. A practical approach for the preparation of high activity Cu/ZnO/ZrO2 catalyst for methanol synthesis from CO2 hydrogenation. Appl. Catal. A Gen. 1998, 171, 45–55. [Google Scholar] [CrossRef]
Figure 1. Influence of carrier composition on the textural properties of Cu-Zn based catalysts.
Figure 1. Influence of carrier composition on the textural properties of Cu-Zn based catalysts.
Chemengineering 01 00019 g001
Figure 2. X-ray diffraction (XRD) patterns of Cu-ZnO based catalysts: (a) as prepared and (b) reduced.
Figure 2. X-ray diffraction (XRD) patterns of Cu-ZnO based catalysts: (a) as prepared and (b) reduced.
Chemengineering 01 00019 g002
Figure 3. Apparent activation energy (Eapp) of Cu–ZnO catalysts at 0.1 MPa.
Figure 3. Apparent activation energy (Eapp) of Cu–ZnO catalysts at 0.1 MPa.
Chemengineering 01 00019 g003
Figure 4. Activity pattern in the CO2 hydrogenation reaction.
Figure 4. Activity pattern in the CO2 hydrogenation reaction.
Chemengineering 01 00019 g004
Figure 5. Structure-reactivity relationships for Cu-ZnO catalysts: (a) Reaction rate as function of MSA at 5.0 MPa; (b) Normalized methanol yield vs. MSA/SAR at 5.0 MPa.
Figure 5. Structure-reactivity relationships for Cu-ZnO catalysts: (a) Reaction rate as function of MSA at 5.0 MPa; (b) Normalized methanol yield vs. MSA/SAR at 5.0 MPa.
Chemengineering 01 00019 g005
Table 1. List of catalytic system based on Cu-Zn oxides composites.
Table 1. List of catalytic system based on Cu-Zn oxides composites.
Sample Chemical Composition (wt%)Physical Properties
CuOZnOZrO2CeO2SA (m2·g−1)PV (cm3·g−1)APD (a) (nm)MSA (c) (m2·g−1)MSA/SAR
CuZn_CZ_100 44.213.442.10.01540.349.0--
R (b)----1320.288.0730.55
CuZn_CZ_95 42.813.238.94.21250.4011.2--
R (b)----620.2012.9280.45
CuZn_CZ_90 42.612.834.08.31140.6018.7--
R (b)----560.2110.8250.44
CuZn_CZ_50 39.012.315.631.8800.8948.6--
R (b)----400.2015.7160.40
CuZn_CZ_20 40.113.05.840.3680.3220.8--
R (b)----360.2021.0140.38
CuZn_CZ_0 38.712.20.048.8470.2420.0--
R (b)----340.1720.0100.29
(a) Average Pore Diameter (APD, 4PV/SA); (b) sample after reduction treatment at 573 K (1 h) in flowing H2/Ar and passivation; (c) obtained by ‘‘single-pulse’’ N2O titration measurements.
Table 2. Results of the catalytic test in the CO2 hydrogenation at different temperature and pressure.
Table 2. Results of the catalytic test in the CO2 hydrogenation at different temperature and pressure.
0.1 MPa X H 2 (%) S CH 3 OH (%)Reaction Rate (mmolH2·h−1·gcat−1)
453 K473 K493 K513 K453 K473 K493 K513 K453 K473 K493 K513 K
CUZN_CZ_1000.51.44.98.934261330.461.374.858.69
CUZN_CZ_950.42.63.66.256452370.372.563.576.04
CUZN_CZ_900.52.04.16.3715230110.461.924.026.22
CUZN_CZ_200.41.22.83.2776239190.371.192.743.11
CUZN_CZ_00.71.01.92.3969280610.731.011.832.29
3.0 MPa X H 2 (%) S CH 3 OH (%)Reaction Rate (mmolH2·h−1·gcat−1)
453 K473 K493 K513 K453 K473 K493 K513 K453 K473 K493 K513 K
CUZN_CZ_1004.17.813.117.7877457504.027.6812.8117.38
CUZN_CZ_953.55.57.713.1938665573.385.407.5912.81
CUZN_CZ_903.55.59.114.9938771503.385.408.8714.63
CUZN_CZ_202.94.26.07.9949082692.844.125.857.77
CUZN_CZ_01.93.24.86.7979590801.833.114.666.59
5.0 MPa X H 2 (%) S CH 3 OH (%)Reaction Rate (mmolH2·h−1·gcat−1)
453 K473 K493 K513 K453 K473 K493 K513 K453 K473 K493 K513 K
CUZN_CZ_1005.39.315.920.5907865645.219.1515.5520.12
CUZN_CZ_954.46.39.314.9938974544.306.229.1514.63
CUZN_CZ_904.16.710.315.9948776564.026.5910.0615.55
CUZN_CZ_203.85.26.99.1949186743.755.126.778.96
CUZN_CZ_02.13.76.17.9979590812.013.665.957.77

Share and Cite

MDPI and ACS Style

Spadaro, L.; Santoro, M.; Palella, A.; Arena, F. Hydrogen Utilization in Green Fuel Synthesis via CO2 Conversion to Methanol over New Cu-Based Catalysts. ChemEngineering 2017, 1, 19. https://doi.org/10.3390/chemengineering1020019

AMA Style

Spadaro L, Santoro M, Palella A, Arena F. Hydrogen Utilization in Green Fuel Synthesis via CO2 Conversion to Methanol over New Cu-Based Catalysts. ChemEngineering. 2017; 1(2):19. https://doi.org/10.3390/chemengineering1020019

Chicago/Turabian Style

Spadaro, Lorenzo, Mariarita Santoro, Alessandra Palella, and Francesco Arena. 2017. "Hydrogen Utilization in Green Fuel Synthesis via CO2 Conversion to Methanol over New Cu-Based Catalysts" ChemEngineering 1, no. 2: 19. https://doi.org/10.3390/chemengineering1020019

Article Metrics

Back to TopTop